Sie sind auf Seite 1von 202

Biodegradable

Metals
Edited by
Eli Aghion
Printed Edition of the Special Issue Published in Metals

www.mdpi.com/journal/metals
Biodegradable Metals
Biodegradable Metals

Special Issue Editor


Eli Aghion

MDPI • Basel • Beijing • Wuhan • Barcelona • Belgrade


Special Issue Editor
Eli Aghion
Ben-Gurion University of the Negev
Israel

Editorial Office
MDPI
St. Alban-Anlage 66
4052 Basel, Switzerland

This is a reprint of articles from the Special Issue published online in the open access journal Metals
(ISSN 2075-4701) from 2017 to 2018 (available at: https://www.mdpi.com/journal/metals/special
issues/biodegradable metals)

For citation purposes, cite each article independently as indicated on the article page online and as
indicated below:

LastName, A.A.; LastName, B.B.; LastName, C.C. Article Title. Journal Name Year, Article Number,
Page Range.

ISBN 978-3-03897-386-7 (Pbk)


ISBN 978-3-03897-387-4 (PDF)

Cover image courtesy of Eli Aghion.


c 2018 by the authors. Articles in this book are Open Access and distributed under the Creative
Commons Attribution (CC BY) license, which allows users to download, copy and build upon
published articles, as long as the author and publisher are properly credited, which ensures maximum
dissemination and a wider impact of our publications.
The book as a whole is distributed by MDPI under the terms and conditions of the Creative Commons
license CC BY-NC-ND.
Contents

About the Special Issue Editor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii

Preface to ”Biodegradable Metals” . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix

Eli Aghion
Biodegradable Metals
Reprinted from: Metals 2018, 8, 804, doi:10.3390/met8100804 . . . . . . . . . . . . . . . . . . . . . 1

Lumei Liu, Juan Wang, Teal Russell, Jagannathan Sankar and Yeoheung Yun
The Biological Responses to Magnesium-Based Biodegradable Medical Devices
Reprinted from: Metals 2017, 7, 514, doi:10.3390/met7110514 . . . . . . . . . . . . . . . . . . . . . 5

Muhammad Imran Rahim, Sami Ullah and Peter P. Mueller


Advances and Challenges of Biodegradable Implant Materials with a Focus on
Magnesium-Alloys and Bacterial Infections
Reprinted from: Metals 2018, 8, 532, doi:10.3390/met8070532 . . . . . . . . . . . . . . . . . . . . . 19

Jakub Tkacz, Karolı́na Slouková, Jozef Minda, Juliána Drábiková, Stanislava Fintová,
Pavel Doležal and Jaromı́r Wasserbauer
Influence of the Composition of the Hank’s Balanced Salt Solution on the Corrosion Behavior
of AZ31 and AZ61 Magnesium Alloys
Reprinted from: Metals 2017, 7, 465, doi:10.3390/met7110465 . . . . . . . . . . . . . . . . . . . . . 33

Matěj Březina, Jozef Minda, Pavel Doležal, Michaela Krystýnová, Stanislava Fintová,
Josef Zapletal, Jaromı́r Wasserbauer and Petr Ptáček
Characterization of Powder Metallurgy Processed Pure Magnesium Materials for
Biomedical Applications
Reprinted from: Metals 2017, 7, 461, doi:10.3390/met7110461 . . . . . . . . . . . . . . . . . . . . . 50

Da-Jun Lin, Fei-Yi Hung, Hung-Pang Lee and Ming-Long Yeh


Development of a Novel Degradation-Controlled Magnesium-Based Regeneration Membrane
for Future Guided Bone Regeneration (GBR) Therapy
Reprinted from: Metals 2017, 7, 481, doi:10.3390/met7110481 . . . . . . . . . . . . . . . . . . . . . 72

Reza Alavi, Adhitya Trenggono, Sébastien Champagne and Hendra Hermawan


Investigation on Mechanical Behavior of Biodegradable Iron Foams under Different
Compression Test Conditions
Reprinted from: Metals 2017, 7, 202, doi:10.3390/met7060202 . . . . . . . . . . . . . . . . . . . . . 90

Mohammad Asgari, Ruiqiang Hang, Chang Wang, Zhentao Yu, Zhiyong Li and Yin Xiao
Biodegradable Metallic Wires in Dental and Orthopedic Applications: A Review
Reprinted from: Metals 2018, 8, 212, doi:10.3390/met8040212 . . . . . . . . . . . . . . . . . . . . . 108

Galit Katarivas Levy, Jeremy Goldman and Eli Aghion


The Prospects of Zinc as a Structural Material for Biodegradable Implants—A Review Paper
Reprinted from: Metals 2017, 7, 402, doi:10.3390/met7100402 . . . . . . . . . . . . . . . . . . . . . 140

Alon Kafri, Shira Ovadia, Jeremy Goldman, Jaroslaw Drelich and Eli Aghion
The Suitability of Zn–1.3%Fe Alloy as a Biodegradable Implant Material
Reprinted from: Metals 2018, 8, 153, doi:10.3390/met8030153 . . . . . . . . . . . . . . . . . . . . . 158

v
Michaela Krystýnová, Pavel Doležal, Stanislava Fintová, Matěj Březina, Josef Zapletal and
Jaromı́r Wasserbauer
Preparation and Characterization of Zinc Materials Prepared by Powder Metallurgy
Reprinted from: Metals 2017, 7, 396, doi:10.3390/met7100396 . . . . . . . . . . . . . . . . . . . . . 173

vi
About the Special Issue Editor
Eli Aghion is a Professor in the Department of Materials Engineering and holder of the Stephen and
Edith Berger Chair in Physical Metallurgy at Ben-Gurion University of the Negev, Beer-Sheva, Israel.
He received his Doctor of Science from the Technion, Israel Institute of Technology and served as
Vice-president and Director of the Magnesium Research Institute at Dead Sea Magnesium, a company
affiliated with Israel Chemicals, Ltd. for over a decade. His research activities are mainly related to (i)
the development of biodegradable metal implants, (ii) assessing the mechano-chemical behavior and
corrosion performance of metals and alloys produced by Additive Manufacturing (AM) technology
(3D printing) and (iii) describing the environmental behavior and properties of light metals (Mg, Al,
Ti) with nano/sub-micron structures.

vii
Preface to ”Biodegradable Metals”
Extensive scientific efforts have been dedicated to the development of biodegradable metal
implants, mainly for orthopedic and cardiovascular applications. This was largely due to the
enhanced mechanical properties of metals, as compared to biodegradable polymers. Obviously, the
main focus of such efforts was directed at metals naturally presenting excellent biocompatibility, such
as Mg, Fe and Zn. However, it soon became evident that these metals possess inherent limitations
when used as structural implants in in vivo conditions. Magnesium-based implants showed
accelerated corrosion rates accompanied by hydrogen gas evolution that can lead to premature loss
of mechanical integrity, separation of tissues and gas embolism. Iron-based implants with acceptable
corrosion rates showed accumulation of voluminous corrosion products that repel neighboring cells
and that did not appear to be metabolized or excreted at appreciable rates. Zinc-based implants have
relatively low mechanical properties and reduced corrosion rates that can provoke fibrous
encapsulation processes which critically limit their biodegradable capabilities. In view of these
limitations, this special issue presents the latest scientific advances in developing biodegradable
metal-based systems with special attention to their safe performance in in vivo conditions.

Eli Aghion
Special Issue Editor

ix
metals
Editorial
Biodegradable Metals
Eli Aghion
Department of Materials Engineering, Ben-Gurion University of the Negev, Beer-Sheva, Israel; egyon@bgu.ac.il

Received: 25 September 2018; Accepted: 2 October 2018; Published: 8 October 2018

1. Introduction and Scope


Over the last two decades, significant scientific efforts have been devoted to developing
biodegradable metal implants for orthopedic and cardiovascular applications, mainly due to their
improved mechanical properties compared to those of biodegradable polymers. Naturally, the main
focus of such efforts was directed at structural metals with the best biocompatibility characteristics,
namely magnesium, iron, and zinc, as well as their alloys. However, it soon became evident that
the use of such metal systems in vivo resulted in major problems, limiting their capabilities to act
as acceptable structural materials for biodegradable implants in practical applications. These hurdles
are exemplified by the accelerated corrosion rate of Mg, accompanied by hydrogen gas evolution
that can lead to premature loss of mechanical integrity of the implant, separation of tissues, and,
in extreme situations, gas embolism [1,2]. In the case of iron-based implants, although iron corrodes
at a reasonable rate, it accumulates a voluminous corrosion product that repels neighboring cells
and biological matrices and does not appear to be either metabolized or excreted at an appreciable
rate [3]. As for zinc-based implants, they possess relatively low mechanical properties in vivo, as well
as reduced corrosion rates, which can provoke fibrous encapsulation processes, which limit their
prospects as practical biodegradable implants [4,5]. As such, efforts have been directed at overcoming
these hurdles.
This Special Issue introduces the latest scientific advances in developing innovative biodegradable
metal implants for various applications and examines the safety of such materials. The collection of
high-quality papers comprising this Special Issue is divided into three parts according to the major
metal matrix component involved, namely Mg-, Fe-, and Zn-based implants.

2. Contributions
The first part of this Special Issue addresses scientific progress made with magnesium-based implants.
The paper by Lumei Liu et al. [6] evaluates the safety of biodegradable Mg-based implants by assessing
their impacts at the cellular/molecular level, including in terms of cell adhesion, signaling, immune
response, and tissue growth, during the degradation process. In addition, they also evaluate the effect
of Mg-based implants on gene expression/protein biosynthesis at the site of implantation, as well
as throughout the body. The outcomes of their study serve as the basis for an innovative prediction method
to assess the safety of magnesium-based implants. Muhammad Imran Rahim et al. [7] concentrated
their efforts on bacterial biofilm infections, as well as bone growth stimulation, given the mechanical
forces imposed by magnesium corrosion products. Their novel model for examining implant-derived
infections suggests that host cell adhesion to implants is important to prevent bacterial invasion of the
exposed host tissue surface, and not, as previously thought, to prevent bacterial adhesion to the implant.
According to their model, they predict that passive antibacterial implant-coating strategies would not
be efficacious in vivo. Jakub Tkacz et al. [8] examined the effects of solution composition and material
surface finish on the corrosion degradation behavior of AZ31 and AZ61 alloys by analyzing the
corrosion products created at the implant surface. Their results reveal differences in the response of the
Mg alloys to the commonly used cell culture medium Hank’s balanced salt solution (HBSS), lacking Mg

Metals 2018, 8, 804; doi:10.3390/met8100804 1 www.mdpi.com/journal/metals


Metals 2018, 8, 804

and Ca ions, and enriched HBSS (HBSS+), containing these ions. Although both alloys exhibited better
corrosion resistance in the enriched HBSS+ solution, AZ61 presented higher values of polarization
resistance than did AZ31 in both corrosion solutions. As for surface finish (i.e., ground vs. polished),
no significant effects were observed during EIS measurements of AZ31 alloy in both HBSS solutions.
By contrast, ground samples of AZ61 in HBSS+ solution displayed Rp values which are higher than
those obtained with polished samples. The effect of processing parameters on mechanical properties
and corrosion degradation of pure Mg in terms of powder metallurgy technology was evaluated by
Matej Brezina et al. [9]. They found that cold-compacted samples were quite brittle, with reduced
strength (up to 50 MPa) and accelerated corrosion degradation, as compared to hot-pressed samples
that yielded much higher strength (up to 250 MPa) and significantly improved corrosion resistance.
Overall, applying temperatures treatments of 300 and 400 ◦ C and high pressures of 300–500 MPa
had a significantly positive influence on material bonding, and mechanical and electrochemical
corrosion properties. A higher compaction temperature of 500 ◦ C had a detrimental effect on material
consolidation processes at compacting pressures above 200 MPa. Apart from using Mg-based implants
as a supporting device for orthopedic and cardiovascular applications, Da-Jun Lin et al. [10] evaluated
the possibility of using Mg-5Zn-0.5Zr alloy for dental-guided bone regeneration. They managed to
develop and integrate an optimized solution of heat-treatment and surface fluoride coating to produce
an Mg-based regeneration membrane. Their heat-treated Mg regeneration membrane ARRm-H380 was
able to provide a proper concentration of Mg ions to accelerate early stage bone growth, encouraging
the formation of more than 80% of the new bone.
The second part of the Special Issue is devoted to innovative efforts aimed at using iron
as a structural material for biodegradable implants. These activities were carried out mainly due to
the inherent limitation of iron, manifested by a reduced corrosion rate in vivo. Although different
methods have been proposed for improving the corrosion rate of iron, including modification of
alloying elements, poly (lactic-co-glycolic acid) infiltration, and coating, Reza Alavi et al. [11] examined
the possibility of using iron-foams as an alternative. The objective of their study was to investigate
the mechanical behavior of iron foams with different cell sizes in various compression tests under
dry and wet conditions and after being subjected to degradation in HBSS. In general, they found
that a wet environment did not significantly change the mechanical behavior of the iron foams,
while degradation processes resulted in reductions of elastic modulus, yield strength, and energy
absorption. Another attempt to overcome the reduced corrosion degradation of iron is introduced in
the review from Mohammad Asgari et al. [12], which notes that sandblasting treatment can increase the
degradation rate of pure Fe in simulated body fluid (SBF). The main reasons for the increased corrosion
rate were changes in surface composition, and the high roughness and density of dislocations.
The third and final section of the Special Issue is dedicated to the prospects of zinc as a promising
alternative to magnesium and iron to overcome the critical limitations of these metals in terms of their
suitability for clinical applications. The review paper by Galit Levy Katarivas et al. [13] indicates that
Zn2+ , the main byproduct of zinc metal corrosion, is highly regulated within physiological systems
and plays a critical role in numerous fundamental cellular processes. However, the use of pure Zn
as a biodegradable metal for most medical device applications is limited due to its insufficient strength,
plasticity, and hardness, as well as its reduced corrosion rate. Although a number of zinc alloys with
relatively improved mechanical properties have been developed, such as Zn–Mg and Zn–Al, they
still do not present satisfactory properties for practical biodegradable implants. Innovative efforts
to develop a new zinc-based alloy were carried out by Alon Kafri et al. [14]. Their research aimed
at evaluating the possibility of using Fe as a relatively cathodic biocompatible alloying element in
zinc that can tune the implant degradation rate via microgalvanic effects. The selected Zn–1.3wt %Fe
alloy composition produced by gravity casting was examined both in vitro and in vivo. The absence
of undesirable systemic effects in terms of gain, subject well-being, and hematological characteristics
(i.e., red blood cell, hemoglobin, and white blood cell levels) of rats during 14 weeks of implantation,
as well as adequate histology results in subcutaneous tissues close to the tested implants, suggests

2
Metals 2018, 8, 804

that the Zn–1.3%Fe alloy can be considered as a potential candidate for biodegradable implants.
Another attempt carried out by Michaela Krystynova et al. [15] aimed at developing zinc implants
via powder metallurgy technology. Their study focused on consolidating zinc powders with two
different particle sizes, 7.5 μm and 150 μm, by cold-pressing, followed by sintering and hot-pressing.
The obtained results showed that the mechanical properties of samples made from 150 μm particle
size powder were better than those prepared with 7.5 μm particle size powder.
In summary, the articles comprising this Special Issue clearly demonstrate the potential of
magnesium-, iron-, and zinc-based systems to be a suitable options as structural materials for
biodegradable metals implants. However, looking ahead, there are still many challenges to be overcome
before these options can be realized. These include: (i) Precisely controlling the degradation kinetics
of biocompatible metals and their corrosion products according to in vivo absorption capabilities,
while maintaining mechanical integrity prior to significant degradation processes; (ii) correlating
the mechanical properties of metal implants according to the required properties of the designated
medical device application; and (iii) conducting long-term clinical trials to obtain full biocompatibility
responses. Considering the importance of realizing these goals, it is reasonable to expect solutions
sooner, rather than later.

References
1. Aghion, E.; Levy, G. The effect of Ca on the in vitro corrosion performance of biodegradable Mg-Nd-Y-Zr
alloy. J. Mater. Sci. 2010, 45, 3096–3101. [CrossRef]
2. Aghion, E.; Levy, G.; Ovadia, S. In vivo behavior of biodegradable Mg-Nd-Y-Zr-Ca alloy. J. Mater. Sci. Mater. Med.
2012, 23, 805–812. [CrossRef] [PubMed]
3. Bowen, P.K.; Shearier, E.R.; Zhao, S.; Guillory, R.J., II; Zhao, F.; Goldman, J.; Drelich, J.W. Biodegradable
Metals for Cardiovascular Stents: From Clinical Concerns to Recent Zn-Alloys. Adv. Healthc. Mater. 2016, 5,
1121–1140. [CrossRef] [PubMed]
4. Bakhsheshi-Rad, H.R.; Hamzah, E.; Low, H.T.; Kasiri-Asgarani, M.; Farahany, S.; Akbari, E.; Cho, M.H. Fabrication
of biodegradable Zn-AlMg alloy: mechanical properties, corrosion behavior, cytotoxicity and antibacterial activities.
Mater Sci Eng C. 2017, 73, 215. [CrossRef] [PubMed]
5. Guillory, R.J.; Bowen, P.K.; Hopkins, S.P.; Shearier, E.R.; Earley, E.J.; Gillette, A.A.; Aghion, E.; Bocks, M.;
Drelich, J.W.; Goldman, J. Corrosion Characteristics Dictate the Long-Term Inflammatory Profile of
Degradable Zinc Arterial Implants. ACS Biomater. Sci. Eng. 2016, 2, 2355–2364. [CrossRef]
6. Liu, L.; Wang, J.; Russell, T.; Sankar, J.; Yun, Y. The Biological Responses to Magnesium-Based Biodegradable
Medical Devices. Metals 2017, 7, 514. [CrossRef]
7. Rahim, M.; Ullah, S.; Mueller, P. Advances and Challenges of Biodegradable Implant Materials with a Focus
on Magnesium-Alloys and Bacterial Infections. Metals 2018, 8, 532. [CrossRef]
8. Tkacz, J.; Slouková, K.; Minda, J.; Drábiková, J.; Fintová, S.; Doležal, P.; Wasserbauer, J. Influence of the
Composition of the Hank’s Balanced Salt Solution on the Corrosion Behavior of AZ31 and AZ61 Magnesium
Alloys. Metals 2017, 7, 465. [CrossRef]
9. Březina, M.; Minda, J.; Doležal, P.; Krystýnová, M.; Fintová, S.; Zapletal, J.; Wasserbauer, J.; Ptáček, P.
Characterization of Powder Metallurgy Processed Pure Magnesium Materials for Biomedical Applications.
Metals 2017, 7, 461. [CrossRef]
10. Lin, D.; Hung, F.; Lee, H.; Yeh, M. Development of a Novel Degradation-Controlled Magnesium-Based
Regeneration Membrane for Future Guided Bone Regeneration (GBR) Therapy. Metals 2017, 7, 481. [CrossRef]
11. Alavi, R.; Trenggono, A.; Champagne, S.; Hermawan, H. Investigation on Mechanical Behavior of Biodegradable
Iron Foams under Different Compression Test Conditions. Metals 2017, 7, 202. [CrossRef]
12. Asgari, M.; Hang, R.; Wang, C.; Yu, Z.; Li, Z.; Xiao, Y. Biodegradable Metallic Wires in Dental and Orthopedic
Applications: A Review. Metals 2018, 8, 212. [CrossRef]
13. Katarivas Levy, G.; Goldman, J.; Aghion, E. The Prospects of Zinc as a Structural Material for Biodegradable
Implants—A Review Paper. Metals 2017, 7, 402. [CrossRef]

3
Metals 2018, 8, 804

14. Kafri, A.; Ovadia, S.; Goldman, J.; Drelich, J.; Aghion, E. The Suitability of Zn–1.3%Fe Alloy as a Biodegradable
Implant Material. Metals 2018, 8, 153. [CrossRef]
15. Krystýnová, M.; Doležal, P.; Fintová, S.; Březina, M.; Zapletal, J.; Wasserbauer, J. Preparation and Characterization of
Zinc Materials Prepared by Powder Metallurgy. Metals 2017, 7, 396. [CrossRef]

© 2018 by the author. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

4
metals
Review
The Biological Responses to Magnesium-Based
Biodegradable Medical Devices
Lumei Liu 1,2 , Juan Wang 3 , Teal Russell 1,2 , Jagannathan Sankar 1 and Yeoheung Yun 1,2, *
1 National Science Foundation-Engineering Research Center for Revolutionizing Metallic Biomaterials,
North Carolina Agricultural and Technical State University, Greensboro, NC 27401, USA;
llumei@aggies.ncat.edu (L.L); tarussel@ncat.edu (T.R.); sankar@ncat.edu (J.S.)
2 FIT BEST Laboratory, Department of Chemical, Biological, and Bioengineering,
North Carolina Agricultural and Technical State University, Greensboro, NC 27401, USA
3 Department of Anesthesiology, School of Medicine, Yale University, New Haven, CT 06519, USA;
juan.wang@yale.edu
* Correspondence: yyun@ncat.edu; Tel.: +1-336-285-3226

Received: 20 September 2017; Accepted: 18 November 2017; Published: 21 November 2017

Abstract: The biocompatibility of Magnesium-based materials (MBMs) is critical to the safety of


biodegradable medical devices. As a promising metallic biomaterial for medical devices, the issue of
greatest concern is devices’ safety as degrading products are possibly interacting with local tissue
during complete degradation. The aim of this review is to summarize the biological responses to
MBMs at the cellular/molecular level, including cell adhesion, transportation signaling, immune
response, and tissue growth during the complex degradation process. We review the influence of
MBMs on gene/protein biosynthesis and expression at the site of implantation, as well as throughout
the body. This paper provides a systematic review of the cellular/molecular behavior of local tissue
on the response to Mg degradation, which may facilitate a better prediction of long-term degradation
and the safe use of magnesium-based implants through metal innovation.

Keywords: biological responses; biocompatibility; biodegradable; magnesium-based materials

1. Introduction
The degradability of magnesium-based materials (MBMs) makes these biomaterials a great choice
for clinical devices, especially for orthopedic and cardiovascular applications. The biocompatibility of
MBM refers to their ability to interact with the body organic tissues without causing an unacceptable
degree of harm. From a biological perspective, human tissue can not only tolerate, but can even
benefit from the interaction with MBM implants by proper responses. On the other hand, the
interaction between MBMs and organic tissue in vivo has also been shown to cause phenomena
that are not observed in vitro. In an aqueous environment, whether that be organic tissue or in vitro
cell culture, Mg reacts with water, generating magnesium hydroxide (Mg(OH)2 ) and molecular
hydrogen (H2 ). The biological responses of Mg-based materials have been studied both in vivo and
in vitro [1–5]. In vivo, MBM implantation results in the formation of gas pockets in tissue containing
different concentrations of H2 , O2 , CO2 , and/or N2 ; a high deposition of calcium phosphate (Ca-P),
which acts as a mineral layer between tissue and MBM implants; and an increase in the local pH
of body fluid [2,6–8]. In contrast, there is no formation of gas pockets in vitro since it is freely
released, while in vivo, the gas pockets are trapped by local tissue. Instead, molecular hydrogen
escapes to the atmosphere, and cell-adhesion behavior on the surface of MBM implants indicates
biocompatibility [1,3]. As a product of MBM corrosion, H2 was also found to be a potential antioxidant
that is involved in cell signaling and has a novel role in preventive and therapeutic applications [9–11].
Furthermore, the Ca-P mineral layer that is associated with magnesium can promote osteoinductivity

Metals 2017, 7, 514; doi:10.3390/met7110514 5 www.mdpi.com/journal/metals


Metals 2017, 7, 514

and osteoconductivity, which aids in the biocompatibility of magnesium alloys as a bone regenerative
material [12]. The increase in pH has a positive correlation in hemoglobin picking up oxygen in the
blood based on the Bohr effect and a negative correlation in cell-mediated bone resorption by rat
osteoclasts in vitro [13,14]. To better understand the biological response to MBMs both in vivo and
in vitro, the mechanism of these phenomena should be investigated on the molecular/cellular level.
Like many non-degradable biomaterials, the surface of MBMs is adhered to via protein integrins
(heterodimeric receptors in the cell membrane) from the extracellular matrix, within nanoseconds
after contact with tissue. Integrins are also involved in intracellular signaling and thus participate
in a diverse range of cell functions [15–17]. For cardiovascular applications, MBMs are subject to
the Vroman effect, which is exhibited by the absorption of blood serum proteins to the biomaterial
surface [18]. However, unlike non-degradable biomaterials, at the time of protein adhesion, Mg reacts
with the aqueous environment to generate hydrogen gas (H2 ) and Mg(OH)2 , thus increasing the
concentration of Mg2+ . It is known that the physiologically active form of Mg2+ serves as a catalyst
for over 300 enzymes, including those for ATP synthesis, as well as those that use other nucleotides
to synthesize DNA and RNA [19]. Both MBMs and permanent biomaterials, such as Titanium-based
alloys, are mixed with biocompatible elements (e.g., rear earth, Nb) [20,21]. However, the biological
responses to the added elements and the molecular mechanisms that need to be determined by in vitro
and in vivo cytotoxicity evaluation.
When MBMs are implanted into the lesion area, a layer of proteins rapidly adsorbs from the blood
(or serum). These proteins effectively translate the structure and composition of the foreign surface
into biological signals. The signals that are generated by the recognition of the foreign MBM implant
are then transmitted from the extracellular environment to the interior of the cell to regulate gene
and protein expression; thus, initiating and mediating cellular behaviors like migration, proliferation,
differentiation, and apoptosis in different cell types [22–24]; in addition to stimulating constructive
responses that favor wound healing and tissue integration. This layer of proteins determines the
activation of the coagulation cascade, complements system, platelets, and immune cells, and guides
their interplay, which results in the formation of a transient provisional matrix and the onset of
an inflammatory response from the immune system [25,26]. Further research should be done on this
protein layer and its expression profile to better understand its involvement in the biological response
to MBMs.
Finally, the immune response leads to an encapsulation of the implants, which also indicates the
growth of tissue. The regular foreign body reaction process of encapsulation includes inflammation,
granulation and regeneration, and fibrosis. It has been shown that Mg2+ on bioceramic surfaces
substantially affects the phenotype of osteogenic cells in vivo and in vitro [27–30]. A number of studies
have demonstrated that Mg2+ plays a critical role in bone remodeling and skeletal development [31].
The mechanism of these phenomena is not yet known, but the function of Mg2+ in protein synthesis
and molecular regulation is a possible explanation. Knowing which genes and proteins are expressed
differently due to the influence of MBM implants and how these molecules are affected will not only
give further insight into the biocompatibility of MBMs, but will also indicate whether MBMs influence
other biological functions involving these proteins. This is of great importance to modern MBM
implant design, which should make full use of these differentially expressed molecules to improve
implant integration [32]. In this article, these molecules from local molecular/cellular response to the
degradation of Mg-based alloys are categorized and reviewed based on their involvement in four
functions: cell adhesion, transportation signaling, immune response, and tissue growth.

2. Degradation of Mg-Based Alloys


The degradation behavior of MBMs has been studied and reviewed [33–36]. The mechanism of
MBMs degradation involves the reaction of magnesium with its aqueous environment, which produces

6
Metals 2017, 7, 514

magnesium hydroxide (Mg(OH)2 ) and hydrogen gas (H2 ). A general summary of the corrosion reaction
kinetics that takes place is given below [34,37]:

2Mg → 2Mg+ + 2e− (anodic reaction) (1)

2H2 O + 2e− → H2 + 2OH− (cathodic reaction) (2)

2Mg2+ + 2H2 O → 2Mg2+ + 2OH− + H2 (chemical reaction) (3)



Mg + 2H2 O → Mg 2+
+ 2OH + H2 (overall reaction) (4)

Mg 2+
+ 2OH → Mg(OH)2 (product formation reaction) (5)

Mg(OH)2 → Mg 2+
+ 2OH (product dissolution reaction) (6)

Mg degradation is a dynamic process, including (1) degradation initiation, (2) degradation


rate, (3) degradation product formation, (4) the composition of degradation products, (5) removal
of the product from flow-induced shear stress, and (6) localized pitting with hydrogen evolution.
This complex process is constantly interacting with local tissue, which involves a typical foreign body
reaction composed of macrophages and foreign body giant cells formation [38]. A local physiological
environment, such as loading and flow affects Mg degradation and finding the most important factors
that influence degradation is the key. These dynamic reactions not only produce corrosion products,
such as solid Mg(OH)2 and H2 gas, but also generate charged molecules that might affect cellular and
molecular responses. For example, it has been studied that responding to different concentrations
of Mg2+ , osteosarcoma (U2OS) cells have different gene expression related to cell growth, apoptosis,
inflammation, and migration [39]. While Mg degrades in the body, the neighboring tissue is expected
to regenerate and sustain normal functions. The active interface between degrading MBMs’ surface
and regenerating local tissue should be monitored and controlled to address the medical concern of
biocompatibility [40].

3. Protein-Mediated Cell Adhesion


The MBM implants enhance the adhesion of surrounding cells that are mediated by proteins in the
extracellular matrix. It is known that cell adhesion and morphology influence their proliferation and
differentiation [41]. The ability of biomaterials to adsorb the proteins from serum in a favorable
conformation determines their ability to support cell adhesion and spreading [42]. The MBMs
have this ability, indicating an important aspect of their relative biocompatibility with adjustable
biodegradation [43]. For example, α5β1- and β1-integrin were found to mediate cell adhesion to
biomaterial surfaces. The expression of α5β1-integrin receptor was increased in human bone-derived
cells (HBDC) responding to Mg2+ -enriched substrates [44]. It has also been shown that the presence of
Mg in bioceramics can significantly increase the expression of β1-, α5β1-, and α3β1-integrins that are
vital for osteoblast activity [44,45]. Mg2+ promotes cell adhesion via 5β1- and β1-integrin-associated
signal transduction pathways, which are involved in the enhanced activation of the key signaling
adaptor protein Shc (Src homology collagen), resulting in the enhanced gene expression of extracellular
matrix proteins [46,47]. In our recent studies, we found that platelets have a different adhesion rate on
different MBMs surfaces in dynamic conditions [5]. The major platelet integrin αIIbβ3 in relation to
MBMs has not been studied. This integrin is required for platelet interactions with proteins in plasma
and the extracellular matrices (ECM) that are essential for platelet adhesion and aggregation during
hemostasis and arterial thrombosis [48].
Surface chemistry modification with Mg2+ also plays an important role in focal adhesion kinase
(FAK; pp125FAK )-mediated signal transduction via cell surface integrin-ECM interaction [44]. It has
been shown that FAK expression is enhanced in osteoblasts growing on Al2 O3 -Mg2+ , suggesting
that tyrosine phosphorylation of signaling proteins was enhanced by binding to Mg2+ -supplemented
bioceramics [44]. In addition to Shc and FAK, other key proteins, such as collagen type 1, vitronectin,

7
Metals 2017, 7, 514

and fibronectin, are also highly expressed by osteoblast cells in the presence of Mg [47]. In vitro,
osteoblastic cells and other cell types have been shown to depend primarily on adsorbed vitronectin or
fibronectin for initial adhesion and spreading on various materials, including tissue culture polystyrene,
titanium, stainless steel, and hydroxyapatite [49–51]. Furthermore, vitronectin and/or fibronectin
have been detected among the proteins adsorbed from whole blood and plasma in vitro and in vivo
by implanted surfaces [52–55]. According to the Vroman effect, under stagnant conditions, initial
protein deposition takes place in this sequence: albumin, globulin, fibrinogen, fibronectin, factor XII,
and HMWK [18]. It has been studied that Mg2+ improves smooth muscle cells adhesion at 10 mM
with certain interaction time. This study revealed some genes that related the influence of Mg2+ to
cell adhesion (SERPINE 1) and inflammation (HMOX1, IL-1β) functions [56]. One exception to the
adhesion-promotion effects of Mg2+ is the rapid formation of hydrogen bubbles that accumulated
next to the MBM surfaces [57], which physically occupy the position for cell attachment [5]. However,
this effect can be moderated by the Ca-P mineral layer coating the surface of MBM implants, which
has been shown to enhance cell attachment and spreading [58]. It has also been demonstrated that
pH-related proteins near isoelectric pH adsorb more on uncharged biomaterial surfaces [59–61]. Thus,
the increasing pH of the surroundings and surface ion change caused by MBM corrosion might
decrease cell adhesion.

4. Transportation Signaling
MBM implants increase the concentration of Mg2+ , which may modify its transportation signaling
pathway between intracellular and extracellular space. Intracellular Mg2+ concentration incorporating
with Mg2+ channels is related to cell growth [62–65]. Mg2+ -related functions in the nucleus and
mitochondria, such as ATP synthesis, will change due to the increased amount of Mg2+ transported by
cell membrane magnesium transporters (Figure 1): transient receptor potential melastatin (TRPM) 6
and 7, SLC41A1, CNNM2, and Claudin-16 and 19 (CLDN 16/19). Calcium homeostasis may also be
altered (Figure 2).
TRPM6 and TRPM7 were characterized as magnesium “gatekeepers” on the cell membrane
that monitor cellular magnesium homeostasis [66]. TRPM7 is responsible for intracellular Mg ion
homeostasis in osteoblast cells and plays an important role in osteoblast proliferation and survival [67].
Thus, tight regulation of magnesium homeostasis is crucial for bone health. Another Mg2+ transporter
is SLC41A1, which was found to be expressed in all of the human tissues tested, but at varying
levels, with the heart and testis having the highest expression of the gene [68]. No explanation of
the expression pattern has been given with regard to Mg2+ -related physiology, though it has been
suggested that SLC41 proteins are likely to be the metazoan equivalent of the Mg transporter E
(MgtE) that is found in bacteria [68]. This will need to be verified using one of the now standard
experimental systems for examining transport, especially in terms of the interface between tissue and
MBM implants. Ancient conserved domain protein 2 (ACDP2) is encoded by CNNM2 and regulates
physiological magnesium homeostasis in humans [69]. It belongs to the ACDP family and is widely
expressed in human tissues, with the highest levels of expression in the brain, kidney, and placenta [70].
Furthermore, studies provide evidence for its involvement in magnesium transport [71,72]. Claudins
allow for Mg2+ transport via the paracellular pathway; that is, that they mediate the transport of
Mg ions through the tight junctions between cells that form an epithelial cell layer. In the claudin
family, Claudin-19, which is encoded by the CLDN19 gene, has been implicated in magnesium
transport [73,74]. Claudin-16 allows the selective re-uptake of Mg2+ in the kidney [75]. Defects in
CLDN16 and CLDN19 can cause primary hypomagnesemia, which is characterized by massive renal
magnesium wasting and hypercalciuria, resulting in nephrocalcinosis and renal failure [76].
Federica I. Wolf et al. suggested that the magnesium-deficient condition led to the increased cells
percentage in the G0/G1-phase and the decreased cells percentage in the S-phase of the cell cycle [77].
Hypermagnesemia is uncommonly reported because the kidney is very efficient in excreting excess
magnesium, thus we believe that patients with renal dysfunctions may not be suitable candidates

8
Metals 2017, 7, 514

for MBM implants. Besides this, hypomagnesemia and increased pH also affect cell morphology.
Echinocytes (red blood cells with a spike-like cell membrane) can be seen with mild hemolysis in
hypomagnesemia and are caused by an increase in pH in vitro [78]. At the site of MBM implantation,
the Mg2+ concentration and pH are increased, and it has not been clearly reported whether MBM
implants will increase the number of echinocytes [79], thus causing acanthocytosis. It seems that host
tissue has regulation on the magnesium transporters overcompensate for the increase in magnesium
ion concentration during the corrosion. However, evidence, such as channels behaviors before, during,
and after Mg-based alloys implantation need to be studied.

Figure 1. An illustration of the main magnesium transporters on the cell membrane.

The layer of Ca-P deposition formed between the host tissue and MBM implants indicates the
transportation of Mg2+ has a tight connection with Ca2+ transportation. TRPM7 by itself appears to be
a Ca2+ channel [80], but in the presence of TRPM6, the affinity series of transported cations places Mg2+
above Ca2+ [67,81]. It has been found that the intestinal absorption and the renal excretion of the two
ions are interdependent [82]. Furthermore, the Ca-P layer is the direct cause of vascular calcification [83].
Studies have shown that magnesium reduces calcification in bovine vascular smooth muscle cells
(BVSMC) in a dose-dependent manner. Higher magnesium levels prevented BVSMC calcification
and inhibited the expression of osteogenic proteins, apoptosis induced by β-glycerophosphate (BGP),
and further progression of already established calcification [84]. It has been demonstrated that Mg2+
interferes with calcium homeostasis and Ca-P deposition in vascular smooth muscle cells (Figure 2) in
the following ways: (1) Mg2+ can stabilize the Ca-P complex and inhibit the apatite transformation from
Ca-P, instead forming more soluble magnesium-substituted whitlockite [85–87]; (2) Mg2+ suppresses
apoptosis resulting in the formation of fewer apoptotic bodies; (3) Mg2+ blocks the entry of Ca2+ into the
cells by being transported into cells as a Ca2+ -channel antagonist [88], and then impedes the formation
of Ca-P particles and Ca-P particle matrix vesicles; (4) Mg2+ enters cells through TRPM7 to balance the
expression of calcification promotors and inhibitors by suppressing the negative effect of Pi (inorganic
phosphate, transported by Pit-1 and Pit-2) on calcification inhibitors (MGP and BMP7) and regressing
the activating effect of phosphate on calcification promotors (RUNX2 and BMP2) [89]; (5) Due to
the effect of Mg2+ on these two calcification promotors, vascular smooth muscle cells are prevented
from undergoing osteogenic differentiation and vascular calcification by the same pathway [84];
and, (6) Mg2+ activates calcium-sensing receptor (CaSR), which inhibits vascular smooth muscle cell
calcification [90,91]. Theoretically, Mg2+ should prevent the formation of the Ca-P layer. In reality,

9
Metals 2017, 7, 514

however, a Ca-P layer is still formed between MBM implants and host tissue and its deposition to
tissue depended on the Mg degradation rate [4].

Figure 2. Mg2+ interferes with calcium homeostasis and Ca-P layer deposition. Abbreviations: Pit,
inorganic phosphate transporter; MGP, matrix Gla protein; BMP, bone morphogenetic protein; RUNX2,
runt-related transcription factor 2.

5. Immune Responses
As an implantation biomaterial, MBMs should induce injury, blood-material interactions,
provisional matrix formation, inflammation, chronic inflammation, granulation tissue development,
foreign body reaction, and fibrosis/fibrous capsule development [38,92–95]. Immune cytokines,
such as IL-4 and IL-13, may be involved to induce monocytes adhesion on MBMs surface and
monocytes/macrophage fusion to form foreign body giant cells [38]. However, because of the
degradability of MBMs, the immune responses are affected by the corrosion products and surface
changes of MBMs. Magnesium ions participate in immune responses in numerous ways: as a cofactor
for immunoglobulin synthesis, C'3 convertase, immune cell adherence, antibody-dependent cytolysis,
IgM-lymphocyte binding, macrophage response to lymphokines, T helper cell-B cell adherence, binding
of substance P to lymphoblasts, and antigen binding to macrophage RNA [96]. As biocompatible
materials, MBMs do not elicit a detrimental immune response. In fact, some of the immunological
responses that are generated by MBMs reflect their beneficial properties.
In one in vitro study, the expression of inflammation-related genes (IL-8, PDGF, TGF-β1, Angio1,
βFGF, VEGF, ET-1, CXCR-1, HIF-1α) was either increased or decreased with different magnesium
ion concentrations [39]. In magnesium-deficient rodents, TNFα, IL-1, and IL-6 are increased in
both the serum and bone marrow microenvironment [97]. Low extracellular magnesium increases
endothelial secretion of growth factors and cytokines, such as interleukin 1 (IL-1), which perpetuates
cell dysfunction and affects smooth muscle cell functions [98]. These factors have important
roles in the immune system. For example, IL-1α and IL-1β are cytokines that participate in the
regulation of immune responses, inflammatory reactions, and hematopoiesis [99]. Interleukin 6
(IL-6), also referred to as B-cell stimulatory factor-2 (BSF-2) and interferon beta-2, is a cytokine
involved in a wide variety of immune functions, such as antibody secretion, acute phase reaction,
and inflammation [100]. Interleukin 8 (IL-8), known as a neutrophil chemotactic factor, is a chemokine
produced by macrophages and other cell types, including epithelial cells, airway smooth muscle
cells [101], and endothelial cells.

10
Metals 2017, 7, 514

The most significant aspect of MBMs that are related to the immune response is hydrogen
gas production [102]. The expression of several pro-inflammatory factors can be decreased by
molecular H2 , including TNF-α, IL-6, IL-1β, CCL2, IL-10, TNF-γ, IL-12, CAM-1 [103], HMGB-1 [104],
PGE2 [105], and nuclear factor-κB (NF-κB) [106]. The design of MBM implants should make use of the
immune response to improve implant integration while avoiding its perpetuation, leading to chronic
inflammation and foreign body reactions, and thus loss of intended function [32].

6. Tissue Growth
MBMs implanted into living tissue initiate host immune responses that reflect the first step
of tissue growth [107] and fibrous encapsulation [38]. There were concerns about tissue damage
because of the evolved hydrogen bubbles and alkalization of solution that are caused by magnesium
degradation [43,108]. In some cases, hydrogen bubbles from a degrading MBM surface can be
accumulated next to the implant and separate tissues and tissue layers, which will delay the healing
of the surgery region and lead to the necrosis of tissues [58]. However, promising studies of
magnesium-based biodegradable materials in vivo have shown that they can enhance new bone
formation in the vicinity of implantation, including the enhanced local formation of the periosteum
and endosteum, two distinct membrane layers that cover the outer and inner surfaces of the bone [109].
MBMs have been shown to be non-toxic and can stimulate bone tissue healing because a high
concentration of magnesium ions can lead to bone cell activation [12]. For cardiovascular tissue growth,
we recently studied Magnesium implantation in arteries both ex vivo and in vivo. Though there are gas
pockets in intima around the implanted Mg wire, the tissue showed normal morphology [4]. A complex
signaling network of growth factors includes epidermal growth factor (EGF), fibroblast growth factor
(FGF), granulocyte macrophage colony stimulating factor (GM-CSF), transform growth factor-β
(TGF-β), vascular endothelial growth factor (VEGF), and platelet derived growth factor (PDGF).
This signaling network controls adhesion, migration, proliferation, and differentiation of fibroblasts,
keratinocytes, and endothelial cells in wound healing [110]. According to Vroman Effect [18], during
the vascular wound healing process, blood proteins will deposit on MBMs surface in a provisional
matrix manner, which provides structural, biochemical, and cellular components to processes wound
healing [38].
Increased expression of collagen I extracellular matrix protein was found in human bone-derived
cells (HBDC) responding to Mg2+ -enriched substrates [44], further suggesting that magnesium
promotes bone growth. In addition to magnesium, studies have shown that the Ca-P layer that
is generated by MBM implants can also promote tissue growth during the biodegradation process
both in vivo and in vitro [12,111]. This layer has been proven to facilitate the differentiation and
proliferation of osteoblastic cells in a Ca-P ratio-dependent manner, indicating that the Ca-P layer
promotes bone formation [112]. There is also Ca-P layer formation due to blood-triggered corrosion of
magnesium alloys [113]. The molecular mechanism of this effect has not been discovered yet; however,
it might be related to the ability of the Ca-P layer to increase cell adhesion and spreading.
There are still some molecules that have not been related to MBM implants that are associated with
tissue growth. For example, Damsky has suggested a role for the integrin molecules α5α1 and α3α1 in
bone formation [114]. It has also been shown that inhibitor of κB kinase–nuclear factor-κB (IKK-NF-κB)
inhibits osteoblastic bone formation by restricting the expression of Fos-related antigen-1 (Fra-1),
an essential transcription factor that is involved in bone matrix formation in vitro and in vivo [115].
Therefore, targeting IKK–NF-κB, α5α1, and α3α1 may help to promote bone formation and treat bone
resorption that occurs due to the inflammatory response after MBM implantation.

7. Systematic Integration
The biodegradation of Mg elicits an increase of Mg2+ , hydrogen gas, and other corrosion products
to homeostasis. The molecules that have been proved or might be related to the responses to these
corrosion products are converged in Table 1. The molecules generally function in cell adhesion,

11
Metals 2017, 7, 514

transportation signaling, immune responses, and tissue growth. The further study of key molecules
that are involved in the in vivo and in vitro response to MBM implants, including their functions and
pathway, are advanced approaches to understand the biocompatibility of MBMs.

Table 1. Molecular factors involved in or possibly related to the response to magnesium-based materials
(MBM) implant corrosion products.

Biological
Mg2+ Ca-P H2
Responses
α5β1-, α3β1-, β1-integrins [44],
Cell Adhesion Shc [46], FAK [44], vitronectin and
fibronectin [47,49], SERPINE 1 [56]
Transportation TRPM6/7 [67,81,82], SLC41A1 [68], CaSR [90,91],
Signaling CLDN16/19 [76], CNNM2 [69] BGP [84]
TNF-α, IL-6, IL-1β,
IL-8, PDGF, TGF-β1, Angio1, βFGF,
CCL2 and IL-10, TNF-γ,
Immune VEGF, ET-1, CXCR-1, HIF-1α [39];
IL-12, CAM-1 [103];
Response HMOX1 [56], IL-1, TNFα, IL-6 [97]; IL-1
HMGB-1 [104];
α and IL-1 β [100]; BSF-2 [100]
PGE2 [105], NF-κB [106]
collagen I extracellular matrix protein [44]; EGF, FGF, GM-CSF, TGF-β, VEGF,
Tissue Growth
PDGF [110]; IKK-NF-κB [110,115]; α5α1 and α3α1 [114]

8. Conclusions
The biocompatibility and degradation properties of Mg alloys make them remarkable implant
materials. The most significant problem with MBMs is the difference in their corrosion behavior
between in vitro and in vivo studies, which reflects the difficulty in predicting the biological responses
of MBMs in the in vitro studies. Another problem is the rapid corrosion of MBMs and the products
generated as a result. Systematically understanding the cellular/molecular responses to MBMs
implants in the aspect of cell adhesion, transportation signaling, immune responses, and tissue growth
are innovative strategies to evaluate their long-term safety for clinical use.

Acknowledgments: This work was supported by NIH NIGMS grant (ISC3GM113728), National Science
Foundation (NSF) EAGER grant (1649243), and Engineering Research Center (ERC) for Revolutionizing Metallic
Biomaterials (NSF-0812348) at North Carolina A & T State University.
Author Contributions: Lumei Liu conceptualized, wrote and edited the paper; Juan Wang, Teal Russell and
Yeoheung Yun revised the paper; Juan Wang gave idea and advice; Jagannathan Sankar and Yeoheung Yun gave
instructions and directions.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Kuhlmann, J.; Bartsch, I.; Willbold, E.; Schuchardt, S.; Holz, O.; Hort, N.; Höche, D.; Heineman, W.R.; Witte, F.
Fast escape of hydrogen from gas cavities around corroding magnesium implants. Acta Biomater. 2013, 9,
8714–8721. [CrossRef] [PubMed]
2. McBride, E.D. Absorbable metal in bone surgery: A further report on the use of magnesium alloys. J. Am.
Med. Assoc. 1938, 111, 2464–2467. [CrossRef]
3. Zhang, S.; Li, J.; Song, Y.; Zhao, C.; Zhang, X.; Xie, C.; Zhang, Y.; Tao, H.; He, Y.; Jiang, Y. In vitro degradation,
hemolysis and mc3t3-e1 cell adhesion of biodegradable Mg–Zn alloy. Mater. Sci. Eng. C 2009, 29, 1907–1912.
[CrossRef]
4. Wang, J.; Liu, L.; Wu, Y.; Maitz, M.F.; Wang, Z.; Koo, Y.; Zhao, A.; Sankar, J.; Kong, D.; Huang, N. Ex vivo
blood vessel bioreactor for analysis of the biodegradation of magnesium stent models with and without
vessel wall integration. Acta Biomater. 2017, 50, 546–555. [CrossRef] [PubMed]

12
Metals 2017, 7, 514

5. Liu, L.; Koo, Y.; Collins, B.; Xu, Z.; Sankar, J.; Yun, Y. Biodegradability and platelets adhesion assessment of
magnesium-based alloys using a microfluidic system. PLoS ONE 2017, 12, e0182914. [CrossRef] [PubMed]
6. McCord, C.P.; Prendergast, J.J.; Meek, S.F.; Harrold, G.C. Chemical gas gangrene from metallic magnesium.
Ind. Med. 1942, 11, 71–75.
7. Edwards, J.D. Application of the Interferometer to Gas Analysis; Government Publishing Office:
Washington, DC, USA, 1919.
8. Song, G.; Song, S.-Z. A possible biodegradable magnesium implant material. Adv. Eng. Mater. 2007, 9,
298–302. [CrossRef]
9. Ohsawa, I.; Ishikawa, M.; Takahashi, K.; Watanabe, M.; Nishimaki, K.; Yamagata, K.; Katsura, K.-I.;
Katayama, Y.; Asoh, S.; Ohta, S. Hydrogen acts as a therapeutic antioxidant by selectively reducing cytotoxic
oxygen radicals. Nat. Med. 2007, 13, 688–694. [CrossRef] [PubMed]
10. Salganik, R.I. The benefits and hazards of antioxidants: Controlling apoptosis and other protective
mechanisms in cancer patients and the human population. J. Am. Coll. Nutr. 2001, 20, 464S–472S. [CrossRef]
[PubMed]
11. Liu, H.; Colavitti, R.; Rovira, I.I.; Finkel, T. Redox-dependent transcriptional regulation. Circ. Res. 2005, 97,
967–974. [CrossRef] [PubMed]
12. Witte, F.; Kaese, V.; Haferkamp, H.; Switzer, E.; Meyer-Lindenberg, A.; Wirth, C.J.; Windhagen, H. In vivo
corrosion of four magnesium alloys and the associated bone response. Biomaterials 2005, 26, 3557–3563.
[CrossRef] [PubMed]
13. Arnett, T.R.; Dempster, D.W. Effect of ph on bone resorption by rat osteoclasts in vitro. Endocrinology 1986,
119, 119–124. [CrossRef] [PubMed]
14. Bohr, C.; Hasselbalch, K.; Krogh, A. Concerning a biologically important relationship—The influence of the
carbon dioxide content of blood on its oxygen binding. Skand. Arch. Physiol. 1904, 16, 402. [CrossRef]
15. Schwartz, M.A.; Schaller, M.D.; Ginsberg, M.H. Integrins: Emerging paradigms of signal transduction.
Annu. Rev. Cell Dev. Biol. 1995, 11, 549–599. [CrossRef] [PubMed]
16. Grzesik, W.J. Integrins and bone—Cell adhesion and beyond. Arch. Immunol. Ther. Exp. 1996, 45, 271–275.
17. Damsky, C.H.; Ilić, D. Integrin signaling: It’s where the action is. Curr. Opin. Cell Biol. 2002, 14, 594–602.
[CrossRef]
18. Vroman, L.; Adams, A.; Fischer, G.; Munoz, P. Interaction of high molecular weight kininogen, factor XII,
and fibrinogen in plasma at interfaces. Blood 1980, 55, 156–159. [PubMed]
19. Coté, C.J.; Lerman, J.; Todres, I.D. A Practice of Anesthesia for Infants and Children E-Book: Expert Consult:
Online and Print; Elsevier Health Sciences: Amsterdam, The Netherlands, 2012.
20. Ehtemam-Haghighi, S.; Liu, Y.; Cao, G.; Zhang, L.-C. Phase transition, microstructural evolution and
mechanical properties of Ti-Nb-Fe alloys induced by fe addition. Mater. Des. 2016, 97, 279–286. [CrossRef]
21. Ehtemam-Haghighi, S.; Prashanth, K.; Attar, H.; Chaubey, A.K.; Cao, G.; Zhang, L. Evaluation of mechanical
and wear properties of Ti-xNb-7Fe alloys designed for biomedical applications. Mater. Des. 2016, 111,
592–599. [CrossRef]
22. Hynes, R.O. Integrins: Versatility, modulation and signaling. Cell 1992, 69, 11–25. [CrossRef]
23. Zhu, X.; Ohtsubo, M.; Böhmer, R.M.; Roberts, J.M.; Assoian, R.K. Adhesion-dependent cell cycle progression
linked to the expression of cyclin D1, activation of cyclin E-cdk2, and phosphorylation of the retinoblastoma
protein. J. Cell Biol. 1996, 133, 391–403. [CrossRef] [PubMed]
24. Van der Flier, A.; Sonnenberg, A. Structural and functional aspects of filamins. Biochim. Biophys. Acta (BBA)
Mol. Cell Res. 2001, 1538, 99–117. [CrossRef]
25. Wilson, C.J.; Clegg, R.E.; Leavesley, D.I.; Pearcy, M.J. Mediation of biomaterial-cell interactions by adsorbed
proteins: A review. Tissue Eng. 2005, 11, 1–18. [CrossRef] [PubMed]
26. Gorbet, M.B.; Sefton, M.V. Biomaterial-associated thrombosis: Roles of coagulation factors, complement,
platelets and leukocytes. Biomaterials 2004, 25, 5681–5703. [CrossRef] [PubMed]
27. Howlett, C.R.; Zreiqat, H.; Wu, Y.; McFall, D.W.; McKenzie, D.R. Effect of ion modification of commonly
used orthopedic materials on the attachment of human bone-derived cells. J. Biomed. Mater. Res. 1999, 45,
345–354. [CrossRef]
28. Zreiqat, H.; Evans, P.; Howlett, C.R. Effect of surface chemical modification of bioceramic on phenotype of
human bone-derived cells. J. Biomed. Mater. Res. 1999, 44, 389–396. [CrossRef]

13
Metals 2017, 7, 514

29. Bilek, M.; Evans, P.; Mckenzie, D.; McCulloch, D.; Zreiqat, H.; Howlett, C. Metal ion implantation using
a filtered cathodic vacuum arc. J. Appl. Phys. 2000, 87, 4198–4204. [CrossRef]
30. Kokubo, T. Bioceramics and Their Clinical Applications; Elsevier: Amsterdam, The Netherlands, 2008.
31. Anast, C.S.; Mohs, J.M.; Kaplan, S.L.; Burns, T.W. Evidence for parathyroid failure in magnesium deficiency.
Science 1972, 177, 606–608. [CrossRef] [PubMed]
32. Williams, D.F. On the mechanisms of biocompatibility. Biomaterials 2008, 29, 2941–2953. [CrossRef] [PubMed]
33. Neil, W.; Forsyth, M.; Howlett, P.; Hutchinson, C.; Hinton, B. Corrosion of magnesium alloy ZE41—The role
of microstructural features. Corros. Sci. 2009, 51, 387–394. [CrossRef]
34. Song, G.L.; Atrens, A. Corrosion mechanisms of magnesium alloys. Adv. Eng. Mater. 1999, 1, 11–33.
[CrossRef]
35. Smith, C.E.; Xu, Z.; Waterman, J.; Sankar, J. Cytocompatibility assessment of mgznca alloys. Emerg. Mater. Res.
2013, 2, 283–290. [CrossRef]
36. Atrens, A.; Liu, M.; Abidin, N.I.Z. Corrosion mechanism applicable to biodegradable magnesium implants.
Mater. Sci. Eng. B 2011, 176, 1609–1636. [CrossRef]
37. Witte, F.; Hort, N.; Vogt, C.; Cohen, S.; Kainer, K.U.; Willumeit, R.; Feyerabend, F. Degradable biomaterials
based on magnesium corrosion. Curr. Opin. Solid State Mater. Sci. 2008, 12, 63–72. [CrossRef]
38. Anderson, J.M.; Rodriguez, A.; Chang, D.T. Foreign body reaction to biomaterials. In Seminars in Immunology;
Elsevier: Amsterdam, The Netherlands, 2008; pp. 86–100.
39. Yun, Y.; Dong, Z.; Tan, Z.; Schulz, M.J. Development of an electrode cell impedance method to measure
osteoblast cell activity in magnesium-conditioned media. Anal. Bioanal. Chem. 2010, 396, 3009–3015.
[CrossRef] [PubMed]
40. Yun, Y.; Dong, Z.; Lee, N.; Liu, Y.; Xue, D.; Guo, X.; Kuhlmann, J.; Doepke, A.; Halsall, H.B.; Heineman, W.
Revolutionizing biodegradable metals. Mater. Today 2009, 12, 22–32. [CrossRef]
41. Ingber, D.E. Tensegrity ii. How structural networks influence cellular information processing networks.
J. Cell Sci. 2003, 116, 1397–1408. [CrossRef] [PubMed]
42. Steele, J.G.; Dalton, B.A.; Johnson, G.; Underwood, P.A. Adsorption of fibronectin and vitronectin onto
primaria™ and tissue culture polystyrene and relationship to the mechanism of initial attachment of human
vein endothelial cells and BHK-21 fibroblasts. Biomaterials 1995, 16, 1057–1067. [CrossRef]
43. Song, G. Control of biodegradation of biocompatable magnesium alloys. Corros. Sci. 2007, 49, 1696–1701.
[CrossRef]
44. Zreiqat, H.; Howlett, C.R.; Zannettino, A.; Evans, P.; Schulze-Tanzil, G.; Knabe, C.; Shakibaei, M. Mechanisms
of magnesium-stimulated adhesion of osteoblastic cells to commonly used orthopaedic implants. J. Biomed.
Mater. Res. 2002, 62, 175–184. [CrossRef] [PubMed]
45. Gronthos, S.; Stewart, K.; Graves, S.E.; Hay, S.; Simmons, P.J. Integrin expression and function on human
osteoblast-like cells. J. Bone Min. Res. 1997, 12, 1189–1197. [CrossRef] [PubMed]
46. Shakibaei, M.; Schulze-Tanzil, G.; de Souza, P.; John, T.; Rahmanzadeh, M.; Rahmanzadeh, R.; Merker, H.-J.
Inhibition of mitogen-activated protein kinase kinase induces apoptosis of human chondrocytes. J. Biol. Chem.
2001, 276, 13289–13294. [CrossRef] [PubMed]
47. Schlaepfer, D.D.; Hanks, S.K.; Hunter, T.; van der Geer, P. Integrin-mediated signal transduction linked to ras
pathway by GRB2 binding to focal adhesion kinase. Nature 1994, 372, 786–791. [CrossRef] [PubMed]
48. Shattil, S.J.; Newman, P.J. Integrins: Dynamic scaffolds for adhesion and signaling in platelets. Blood 2004,
104, 1606–1615. [CrossRef] [PubMed]
49. Steele, J.G.; McFarland, C.; Dalton, B.A.; Johnson, G.; Evans, M.D.; Rolfe Howlett, C.; Underwood, P.A.
Attachment of human bone cells to tissue culture polystyrene and to unmodified polystyrene: The effect of
surface chemistry upon initial cell attachment. J. Biomater. Sci. Polym. Ed. 1994, 5, 245–257. [CrossRef]
50. Howlett, C.R.; Evans, M.D.; Walsh, W.R.; Johnson, G.; Steele, J.G. Mechanism of initial attachment of cells
derived from human bone to commonly used prosthetic materials during cell culture. Biomaterials 1994, 15,
213–222. [CrossRef]
51. Kilpadi, K.L.; Chang, P.L.; Bellis, S.L. Hydroxylapatite binds more serum proteins, purified integrins,
and osteoblast precursor cells than titanium or steel. J. Biomed. Mater. Res. 2001, 57, 258–267. [CrossRef]
52. Bale, M.D.; Wohlfahrt, L.A.; Mosher, D.F.; Tomasini, B.; Sutton, R.C. Identification of vitronectin as a major
plasma protein adsorbed on polymer surfaces of different copolymer composition. Blood 1989, 74, 2698–2706.
[PubMed]

14
Metals 2017, 7, 514

53. Fabrizius-Homan, D.J.; Cooper, S.L. A comparison of the adsorption of three adhesive proteins to biomaterial
surfaces. J. Biomater. Sci. Polym. Ed. 1992, 3, 27–47. [CrossRef]
54. Babensee, J.E.; Cornelius, R.M.; Brash, J.L.; Sefton, M.V. Immunoblot analysis of proteins associated
with hema-mma microcapsules: Human serum proteins in vitro and rat proteins following implantation.
Biomaterials 1998, 19, 839–849. [CrossRef]
55. Rosengren, Å.; Pavlovic, E.; Oscarsson, S.; Krajewski, A.; Ravaglioli, A.; Piancastelli, A. Plasma protein
adsorption pattern on characterized ceramic biomaterials. Biomaterials 2002, 23, 1237–1247. [CrossRef]
56. Ma, J.; Zhao, N.; Zhu, D. Biphasic responses of human vascular smooth muscle cells to magnesium ion.
J. Biomed. Mater. Res. Part A 2016, 104, 347–356. [CrossRef] [PubMed]
57. Meyer-Lindenberg, A.; Windhugen, H.; Witte, F. Medical Implant for the Human or Animal Body.
U.S. Patents US20,040,241,036 A1, 2 December 2004.
58. Maxian, S.H.; Zawadsky, J.P.; Dunn, M.G. Effect of Ca/P coating resorption and surgical fit on the
bone/implant interface. J. Biomed. Mater. Res. 1994, 28, 1311–1319. [CrossRef] [PubMed]
59. Norde, W. Driving forces for protein adsorption at solid surfaces. In Macromolecular Symposia; Wiley Online
Library: New York, NY, USA, 1996; pp. 5–18.
60. Ohno, Y.; Maehashi, K.; Yamashiro, Y.; Matsumoto, K. Electrolyte-gated graphene field-effect transistors for
detecting pH and protein adsorption. Nano Lett. 2009, 9, 3318–3322. [CrossRef] [PubMed]
61. Dee, K.C.; Puleo, D.A.; Bizios, R. An Introduction to Tissue-Biomaterial Interactions; John Wiley & Sons:
Hoboken, NJ, USA, 2003.
62. Maier, J.A.; Bernardini, D.; Rayssiguier, Y.; Mazur, A. High concentrations of magnesium modulate vascular
endothelial cell behaviour in vitro. Biochim. Biophys. Acta (BBA) Mol. Basis Dis. 2004, 1689, 6–12. [CrossRef]
[PubMed]
63. Moomaw, A.S.; Maguire, M.E. The unique nature of Mg2+ channels. Physiology 2008, 23, 275–285. [CrossRef]
[PubMed]
64. Günther, T. Concentration, compartmentation and metabolic function of intracellular free Mg2+ . Magnes. Res.
2006, 19, 225–236. [PubMed]
65. Bo, S.; Pisu, E. Role of dietary magnesium in cardiovascular disease prevention, insulin sensitivity and
diabetes. Curr. Opin. Lipidol. 2008, 19, 50–56. [CrossRef] [PubMed]
66. Schlingmann, K.P.; Waldegger, S.; Konrad, M.; Chubanov, V.; Gudermann, T. TRPM6 and
TRPM7—Gatekeepers of human magnesium metabolism. Biochim. Biophys. Acta (BBA) Mol. Basis Dis.
2007, 1772, 813–821. [CrossRef] [PubMed]
67. Nadler, M.J.; Hermosura, M.C.; Inabe, K.; Perraud, A.-L.; Zhu, Q.; Stokes, A.J.; Kurosaki, T.; Kinet, J.-P.;
Penner, R.; Scharenberg, A.M. Ltrpc7 is a Mg·ATP-regulated divalent cation channel required for cell viability.
Nature 2001, 411, 590–595. [CrossRef] [PubMed]
68. Wabakken, T.; Rian, E.; Kveine, M.; Aasheim, H.-C. The human solute carrier SLC41A1 belongs to a novel
eukaryotic subfamily with homology to prokaryotic MgtE Mg2+ transporters. Biochem. Biophys. Res. Commun.
2003, 306, 718–724. [CrossRef]
69. Meyer, T.E.; Verwoert, G.C.; Hwang, S.-J.; Glazer, N.L.; Smith, A.V.; van Rooij, F.J.; Ehret, G.B.; Boerwinkle, E.;
Felix, J.F.; Leak, T.S.; et al. Genome-wide association studies of serum magnesium, potassium, and sodium
concentrations identify six loci influencing serum magnesium levels. PLoS Genet. 2010, 6, e1001045.
[CrossRef] [PubMed]
70. Wang, C.-Y.; Shi, J.-D.; Yang, P.; Kumar, P.G.; Li, Q.-Z.; Run, Q.-G.; Su, Y.-C.; Scott, H.S.; Kao, K.-J.; She, J.-X.
Molecular cloning and characterization of a novel gene family of four ancient conserved domain proteins
(ACDP). Gene 2003, 306, 37–44. [CrossRef]
71. Goytain, A.; Quamme, G.A. Functional characterization of ACDP2 (ancient conserved domain protein),
a divalent metal transporter. Physiol. Genom. 2005, 22, 382–389. [CrossRef] [PubMed]
72. Will, C.; Breiderhoff, T.; Thumfart, J.; Stuiver, M.; Kopplin, K.; Sommer, K.; Günzel, D.; Querfeld, U.; Meij, I.C.;
Shan, Q.; et al. Targeted deletion of murine cldn16 identifies extra-and intrarenal compensatory mechanisms
of Ca2+ and Mg2+ wasting. Am. J. Physiol.-Ren. Physiol. 2010, 298, F1152–F1161. [CrossRef] [PubMed]
73. Naeem, M.; Hussain, S.; Akhtar, N. Mutation in the tight-junction gene claudin 19 (CLDN19) and familial
hypomagnesemia, hypercalciuria, nephrocalcinosis (FHHNC) and severe ocular disease. Am. J. Nephrol.
2011, 34, 241–248. [CrossRef] [PubMed]

15
Metals 2017, 7, 514

74. Konrad, M.; Schaller, A.; Seelow, D.; Pandey, A.V.; Waldegger, S.; Lesslauer, A.; Vitzthum, H.; Suzuki, Y.;
Luk, J.M.; Becker, C. Mutations in the tight-junction gene claudin 19 (CLDN19) are associated with renal
magnesium wasting, renal failure, and severe ocular involvement. Am. J. Hum. Genet. 2006, 79, 949–957.
[CrossRef] [PubMed]
75. Hou, J.; Renigunta, A.; Konrad, M.; Gomes, A.S.; Schneeberger, E.E.; Paul, D.L.; Waldegger, S.;
Goodenough, D.A. Claudin-16 and claudin-19 interact and form a cation-selective tight junction complex.
J. Clin. Investig. 2008, 118, 619. [CrossRef] [PubMed]
76. Hou, J.; Renigunta, A.; Gomes, A.S.; Hou, M.; Paul, D.L.; Waldegger, S.; Goodenough, D.A. Claudin-16
and claudin-19 interaction is required for their assembly into tight junctions and for renal reabsorption of
magnesium. Proc. Natl. Acad. Sci. USA 2009, 106, 15350–15355. [CrossRef] [PubMed]
77. Wolf, F.I.; Cittadini, A. Magnesium in cell proliferation and differentiation. Front. Biosci. 1999, 4, D607–D617.
[CrossRef] [PubMed]
78. Hoffman, R.; Benz, E.J., Jr.; Silberstein, L.E.; Heslop, H.; Weitz, J.; Anastasi, J. Hematology: Basic Principles
and Practice, Expert Consult Premium Edition-Enhanced Online Features; Elsevier Health Sciences: Amsterdam,
The Netherland, 2012.
79. Wang, J.; He, Y.; Maitz, M.F.; Collins, B.; Xiong, K.; Guo, L.; Yun, Y.; Wan, G.; Huang, N. A surface-eroding poly
(1,3-trimethylene carbonate) coating for fully biodegradable magnesium-based stent applications: Toward
better biofunction, biodegradation and biocompatibility. Acta Biomater. 2013, 9, 8678–8689. [CrossRef]
[PubMed]
80. Runnels, L.W.; Yue, L.; Clapham, D.E. The TRPM7 channel is inactivated by PIP2 hydrolysis. Nat. Cell Biol.
2002, 4, 329–336. [CrossRef] [PubMed]
81. Monteilh-Zoller, M.K.; Hermosura, M.C.; Nadler, M.J.; Scharenberg, A.M.; Penner, R.; Fleig, A.
TRPM7 provides an ion channel mechanism for cellular entry of trace metal ions. J. Gen. Physiol. 2003, 121,
49–60. [CrossRef] [PubMed]
82. Paunier, L. Effect of magnesium on phosphorus and calcium metabolism. Monatsschr. Kinderheilkd. Organ
Deutsch. Ges. Kinderheilkd. 1992, 140, S17–S20.
83. Massy, Z.A.; Drüeke, T.B. Magnesium and cardiovascular complications of chronic kidney disease.
Nat. Rev. Nephrol. 2015, 11, 432–442. [CrossRef] [PubMed]
84. Kircelli, F.; Peter, M.E.; Ok, E.S.; Celenk, F.G.; Yilmaz, M.; Steppan, S.; Asci, G.; Ok, E.; Passlick-Deetjen, J.
Magnesium reduces calcification in bovine vascular smooth muscle cells in a dose-dependent manner.
Nephrol. Dial. Transplant. 2011, 27, 514–521. [CrossRef] [PubMed]
85. LeGeros, R. Formation and transformation of calcium phosphates: Relevance to vascular calcification.
Z. Kardiol. 2001, 90, 116–124. [CrossRef] [PubMed]
86. Cheng, P.-T.; Grabher, J.; LeGeros, R. Effects of magnesium on calcium phosphate formation. Magnesium
1987, 7, 123–132.
87. Peters, F.; Epple, M. Simulating arterial wall calcification in vitro: Biomimetic crystallization of calcium
phosphates under controlled conditions. Z. Kardiol. 2001, 90, 81–85. [CrossRef] [PubMed]
88. Altura, B.; Altura, B.; Carella, A.; Gebrewold, A.; Murakawa, T.; Nishio, A. Mg2+ -Ca2+ interaction in
contractility of vascular smooth muscle: Mg2+ versus organic calcium channel blockers on myogenic tone
and agonist-induced responsiveness of blood vessels. Can. J. Physiol. Pharmacol. 1987, 65, 729–745. [CrossRef]
[PubMed]
89. Montezano, A.C.; Zimmerman, D.; Yusuf, H.; Burger, D.; Chignalia, A.Z.; Wadhera, V.; van Leeuwen, F.N.;
Touyz, R.M. Vascular smooth muscle cell differentiation to an osteogenic phenotype involves TRPM7
modulation by magnesium. Hypertension 2010, 56, 453–462. [CrossRef] [PubMed]
90. Brown, E.M.; Gamba, G.; Riccardi, D.; Lombardi, M.; Butters, R.; Kifor, O.; Sun, A.; Hediger, M.A.; Lytton, J.;
Hebert, S.C. Cloning and characterization of an extracellular Ca2+ -sensing receptor from bovine parathyroid.
Nature 1993, 366, 575–580. [CrossRef] [PubMed]
91. Ivanovski, O.; Nikolov, I.G.; Joki, N.; Caudrillier, A.; Phan, O.; Mentaverri, R.; Maizel, J.; Hamada, Y.;
Nguyen-Khoa, T.; Fukagawa, M. The calcimimetic R-568 retards uremia-enhanced vascular calcification and
atherosclerosis in apolipoprotein E deficient (apoE−/−) mice. Atherosclerosis 2009, 205, 55–62. [CrossRef]
[PubMed]
92. Anderson, J.M. Biological responses to materials. Annu. Rev. Mater. Res. 2001, 31, 81–110. [CrossRef]
93. Anderson, J.M. Multinucleated giant cells. Curr. Opin. Hematol. 2000, 7, 40–47. [CrossRef] [PubMed]

16
Metals 2017, 7, 514

94. Gretzer, C.; Emanuelsson, L.; Liljensten, E.; Thomsen, P. The inflammatory cell influx and cytokines changes
during transition from acute inflammation to fibrous repair around implanted materials. J. Biomater. Sci.
Polym. Ed. 2006, 17, 669–687. [CrossRef] [PubMed]
95. Luttikhuizen, D.T.; Harmsen, M.C.; Luyn, M.J.V. Cellular and molecular dynamics in the foreign body
reaction. Tissue Eng. 2006, 12, 1955–1970. [CrossRef] [PubMed]
96. Galland, L. Magnesium and immune function: An overview. Magnesium 1988, 7, 290–299. [PubMed]
97. Mazur, A.; Maier, J.A.M.; Rock, E.; Gueux, E.; Nowacki, W.; Rayssiguier, Y. Magnesium and the inflammatory
response: Potential physiopathological implications. Arch. Biochem. Biophys. 2007, 458, 48–56. [CrossRef]
[PubMed]
98. Maier, J.A.; Malpuech-Brugère, C.; Zimowska, W.; Rayssiguier, Y.; Mazur, A. Low magnesium promotes
endothelial cell dysfunction: Implications for atherosclerosis, inflammation and thrombosis. Biochim. Biophys.
Acta (BBA) Mol. Basis Dis. 2004, 1689, 13–21. [CrossRef] [PubMed]
99. Sims, J.E.; March, C.J.; Cosman, D.; Widmer, M.B.; MacDonald, H.R.; McMahan, C.J.; Grubin, C.E.;
Wignall, J.M.; Jackson, J.L.; Call, S.M.; et al. cDNA expression cloning of the IL-1 receptor, a member
of the immunoglobulin superfamily. Science 1988, 241, 585–589. [CrossRef] [PubMed]
100. Hirano, T.; Yasukawa, K.; Harada, H.; Taga, T.; Watanabe, Y.; Matsuda, T.; Kashiwamura, S.-I.; Nakajima, K.;
Koyama, K.; Iwamatsu, A.; et al. Complementary DNA for a novel human interleukin (BSF-2) that induces b
lymphocytes to produce immunoglobulin. Nature 1986, 324, 73–76. [CrossRef] [PubMed]
101. Hedges, J.C.; Singer, C.A.; Gerthoffer, W.T. Mitogen-activated protein kinases regulate cytokine gene
expression in human airway myocytes. Am. J. Respir. Cell Mol. Biol. 2000, 23, 86–94. [CrossRef] [PubMed]
102. Seitz, J.M.; Eifler, R.; Bach, F.; Maier, H.J. Magnesium degradation products: Effects on tissue and human
metabolism. J. Biomed. Mater. Res. Part A 2014, 102, 3744–3753. [CrossRef] [PubMed]
103. Buchholz, B.M.; Kaczorowski, D.J.; Sugimoto, R.; Yang, R.; Wang, Y.; Billiar, T.R.; McCurry, K.R.; Bauer, A.J.;
Nakao, A. Hydrogen inhalation ameliorates oxidative stress in transplantation induced intestinal graft injury.
Am. J. Transplant. 2008, 8, 2015–2024. [CrossRef] [PubMed]
104. Xie, K.; Yu, Y.; Pei, Y.; Hou, L.; Chen, S.; Xiong, L.; Wang, G. Protective effects of hydrogen gas on murine
polymicrobial sepsis via reducing oxidative stress and hmgb1 release. Shock 2010, 34, 90–97. [CrossRef]
[PubMed]
105. Kawasaki, H.; Guan, J.; Tamama, K. Hydrogen gas treatment prolongs replicative lifespan of bone
marrow multipotential stromal cells in vitro while preserving differentiation and paracrine potentials.
Biochem. Biophys. Res. Commun. 2010, 397, 608–613. [CrossRef] [PubMed]
106. Chen, H.; Sun, Y.P.; Li, Y.; Liu, W.W.; Xiang, H.G.; Fan, L.Y.; Sun, Q.; Xu, X.Y.; Cai, J.M.; Ruan, C.P.;
et al. Hydrogen-rich saline ameliorates the severity of L-arginine-induced acute pancreatitis in rats.
Biochem. Biophys. Res. Commun. 2010, 393, 308–313. [CrossRef] [PubMed]
107. Franz, S.; Rammelt, S.; Scharnweber, D.; Simon, J.C. Immune responses to implants—A review of the
implications for the design of immunomodulatory biomaterials. Biomaterials 2011, 32, 6692–6709. [CrossRef]
[PubMed]
108. Song, G. Recent progress in corrosion and protection of magnesium alloys. Adv. Eng. Mater. 2005, 7, 563–586.
[CrossRef]
109. Zhang, E.; Xu, L.; Yu, G.; Pan, F.; Yang, K. In vivo evaluation of biodegradable magnesium alloy bone implant
in the first 6 months implantation. J. Biomed. Mater. Res. Part A 2009, 90, 882–893. [CrossRef] [PubMed]
110. Barrientos, S.; Stojadinovic, O.; Golinko, M.S.; Brem, H.; Tomic-Canic, M. Growth factors and cytokines in
wound healing. Wound Repair Regen. 2008, 16, 585–601. [CrossRef] [PubMed]
111. Li, L.; Gao, J.; Wang, Y. Evaluation of cyto-toxicity and corrosion behavior of alkali-heat-treated magnesium
in simulated body fluid. Surf. Coat. Technol. 2004, 185, 92–98. [CrossRef]
112. Hulshoff, J.; Van Dijk, K.; De Ruijter, J.; Rietveld, F.; Ginsel, L.; Jansen, J. Interfacial phenomena: An in vitro
study of the effect of calcium phosphate (Ca-P) ceramic on bone formation. J. Biomed. Mater. Res. 1998, 40,
464–474. [CrossRef]
113. Geis-Gerstorfer, J.; Schille, C.; Schweizer, E.; Rupp, F.; Scheideler, L.; Reichel, H.P.; Hort, N.; Nolte, A.;
Wendel, H.P. Blood triggered corrosion of magnesium alloys. Mater. Sci. Eng. B 2011, 176, 1761–1766.
[CrossRef]

17
Metals 2017, 7, 514

114. Damsky, C.H. Extracellular matrix-integrin interactions in osteoblast function and tissue remodeling. Bone
1999, 25, 95–96. [CrossRef]
115. Chang, J.; Wang, Z.; Tang, E.; Fan, Z.; McCauley, L.; Franceschi, R.; Guan, K.; Krebsbach, P.H.; Wang, C.-Y.
Inhibition of osteoblastic bone formation by nuclear factor-κB. Nat. Med. 2009, 15, 682–689. [CrossRef]
[PubMed]

© 2017 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

18
metals
Review
Advances and Challenges of Biodegradable Implant
Materials with a Focus on Magnesium-Alloys and
Bacterial Infections
Muhammad Imran Rahim 1 , Sami Ullah 2 and Peter P. Mueller 3, *
1 Department of Prosthetic Dentistry and Biomedical Materials Science, Lower Saxony Centre for Biomedical
Engineering, Implant Research and Development, Hannover Medical School, Carl-Neuberg-Straße 1,
30625 Hannover, Germany; M.Imran.Rahim@outlook.com
2 Department of MSYS, Helmholtz Center for Infection Research, Inhoffenstrasse 7,
38124 Braunschweig, Germany; Sami.Ullah@helmholtz-hzi.de
3 Department of Chemical Biology, Helmholtz Center for Infection Research, Inhoffenstrasse 7,
38124 Braunschweig, Germany
* Correspondence: pmu@gbf.de; Tel.: +49-531-6181-5070

Received: 12 June 2018; Accepted: 4 July 2018; Published: 10 July 2018

Abstract: Medical implants made of biodegradable materials could be advantageous for temporary
applications, such as mechanical support during bone-healing or as vascular stents to keep blood
vessels open. After completion of the healing process, the implant would disappear, avoiding
long-term side effects or the need for surgical removal. Various corrodible metal alloys based
on magnesium, iron or zinc have been proposed as sturdier and potentially less inflammatory
alternatives to degradable organic polymers, in particular for load-bearing applications. Despite the
recent introduction of magnesium-based screws, the remaining hurdles to routine clinical applications
are still challenging. These include limitations such as mechanical material characteristics or
unsuitable corrosion characteristics. In this article, the salient features and clinical prospects of
currently-investigated biodegradable implant materials are summarized, with a main focus on
magnesium alloys. A mechanism of action for the stimulation of bone growth due to the exertion
of mechanical force by magnesium corrosion products is discussed. To explain divergent in vitro
and in vivo effects of magnesium, a novel model for bacterial biofilm infections is proposed which
predicts crucial consequences for antibacterial implant strategies.

Keywords: bioresorbable implants; corrosion layer; vascular stents; orthopedic implants;


microbial infections

1. Introduction
For metallic implants, industrially-developed, inert and long-lasting materials, such as titanium
(Ti) alloys, stainless steel (SS) and cobalt–chromium (CoCr) alloys, are most frequently used [1–4].
The duration of the healing process is highly variable depending on the extent of injury, disease
state, age and treatment. In general, healing time may range from a brief one month period, up to
a six month period in more complex cases. Permanent implants are frequently removed after the
completion of the healing process to avoid diverse side effects. Long-term disadvantages of this practice
include the failure to adapt to rapid growth in young children, bone degradation by stress shielding,
microbial implant infections, excessive fibrosis or persistent inflammation. Novel bioresorbable
metal implants could provide support during the healing process, and then disappear to avoid
long-term side effects without requiring surgical removal [5,6]. Conventionally, initial material tests
are done most economically in vitro under precisely defined technical conditions. Subsequent assays
are performed under increasingly complex and more costly cell culture conditions, followed by

Metals 2018, 8, 532; doi:10.3390/met8070532 19 www.mdpi.com/journal/metals


Metals 2018, 8, 532

small animal experiments and eventually tested in large animals or in clinical trials. However,
corrosion results obtained under simple technical conditions cannot be extrapolated to clinical
circumstances. In one study, the corrosion rate for magnesium alloys was reported to differ four orders
of magnitude between in vitro and in vivo conditions [7]. Due to inherent limitations in reproducing
the complexities of living tissue in vitro, this review preferentially refers to animal models or clinical
trials if available [8,9]. For molecular genetic and economic reasons, small animal studies are most
popular. However, for load bearing applications in particular, size is an important parameter that must
be kept in mind, and eventually large animal experiments and clinical data are essential. Even though
intense research efforts recently culminated in clinical reports, degradable metallic implants are not
yet routinely applied (see Section 7 for details). Compared to organic polymers, biodegradable metals
can achieve higher strengths and ductility and would therefore would be preferential for load bearing
applications such as bone plates, screws or coronary stents [10]. Three types of metal alloys have been
commonly-investigated as degradable implants for biomedical applications based on magnesium, iron
or zinc. The main purpose of this article is a brief and easily understandable overview of virtues and
clinical hurdles of self-degrading implants as screws, plates or intramedullary rods for load-bearing
orthopedic (musculoskeletal) applications or as vascular stents. In addition, a novel model for implant
infections is proposed to explain divergent effects of magnesium on bacteria in vitro and in vivo.

2. Material Requirements for Fully-Bioresorbable Vascular Stents


Clinical requirements provide the basis for the required implant material characteristics.
Age-related vascular malfunctions such as vessels clogged by a blood clot are of growing
importance [11,12]. One of the earliest effective treatments was antithrombotic therapy, but this
required some time until the clot was dissolved. Then, vascular stents, used to keep blood vessels
open, were shown to be superior, despite the fact that treatment had to be delayed to prevent patient
deaths. Balloon angioplasty is routinely applied and requires minimal invasive surgery. A small
folded stent on a balloon at the tip of a catheter is maneuvered through blood vessels until it is
located at the site of the restriction. The position within the body can be monitored with the help of
an X-ray camera. For this reason, an X-ray dense stent material is an advantage. Once positioned,
the balloon is inflated to unfold the stent. Stents are overextended to a limited degree to allow a
firm anchoring in the vessel wall to prevent migration, and to compensate for the inherent elastic
recoil of the stent after the balloon had deflated. The stent material must be sturdy enough to allow
for thin struts, minimize the recoil, and withstand the pressure of the tissue and the forces during
movements of the body [13]. It is clinically well established that thin, yet robust and highly ductile,
stainless steel or shape memory alloy stents fulfill these requirements. Nevertheless, initial stent
overextension and subsequent persistent mechanical stress, due to interactions with pulsing blood
vessel walls, stimulates smooth muscle cell proliferation in the vessel walls. In a process, termed
restenosis, a growing mass of proliferating vascular smooth muscle-related cells eventually obstruct
the stented vessel again. The blood flow may be reestablished by inserting a second stent or through
surgical bypass, leading to additional patient discomfort, risks and costs [14,15]. In clinical applications
restenosis has been successfully curbed by drug-eluting stents [16]. Thereby, unwanted cell growth is
suppressed locally by clinically well-established drug-loaded polymer coated stents that gradually
release immune suppressive agents like sirolimus or the antiproliferative-acting drug paclitaxel.
Even though these drugs could reduce the incidence of restenosis, serious side effects include a
delayed healing response, inflammation and persistent thrombosis risks [17–20]. This results in the
requirement of costly regular and prolonged antiplatelet treatments, and non-complying patients
drastically increase the thrombosis hazard [21]. Therefore, the application of drug-eluting stents
must be carefully considered for each patient individually depending on the restenosis risks and the
treatment-associated bleeding vulnerability. As an alternative, long-term side effects could be avoided
by using fully biodegradable stents which provide the essential support for a few weeks during the
healing process, and then completely disappear [22–25]. Key degradable stent material requirements

20
Metals 2018, 8, 532

are appropriate corrosion characteristics, biocompatibility, high elasticity to allow for small folded
stents and sufficient strength to resist collapsing.

3. Material Requirements for Degradable Orthopedic Implants


To allow healing, broken bones must be firmly stabilized to avoid even micro-movements under
the influence of considerable forces. Since inflammation may antagonize bone repair, the implant
must be highly biocompatible. Clinically, all requirements are met by sturdy plates, screws or
intramedullary nails made of titanium alloys or stainless steel. Nevertheless, after completion of
the healing process stress shielding implants are mostly removed since their prolonged presence
can lead to bone degradation [26]. Strong, tissue friendly self-degrading implants with bone-like
mechanical parameters, to minimize stress-shielding, and suitable degradation characteristics could
reduce such side effects. Furthermore, they allow patients to avoid a second surgery for implant
removal, making them a highly attractive option. Whereas conventional permanent implant materials
are sturdy and biologically inert, resorbable polymeric materials, as well as corrodible metals, have
distinct biological characteristics (Table 1). In the following table, the cardinal properties of the most
intensively investigated prospective biodegradable implant materials for load-bearing applications
are described.

Table 1. Basic properties of degradable implant materials.

Implant Physical and Corrosion


Degradation Speed Biological Effects References
Material Characteristics
Potentially flexible but mostly
too weak for load-bearing
Organic Inflammatory acidic
Adjustable applications; Implant swelling [27,28]
polymers hydrolysis products
in moist environments; X-ray
transparent
Accumulation of
Very slow, complete
Sturdy but irregular corrosion inflammatory iron
Iron degradation may require [29–31]
characteristics hydroxide particles in
several years
various tissues
Slow, life-time by far
Zinc-based exceeds expected Suboptimal strength Non-inflammatory [32,33]
healing periods
Alloys with sufficient strength
Non-inflammatory; gas
available; compliance can be
accumulation in the
Rapid, danger of adjusted; irregular pitting
tissue; accumulating
mechanical implant corrosion; corrosion coat
Magnesium-based solid corrosion products [34–36]
failure before the healing formation due to slowly
or gaseous hydrogen
process is completed dissolving solid precipitates
may exert pressure on
resulting in reduction of initial
non-yielding bony tissue
corrosion rates
Sturdy, suitable for
load-bearing applications,
Surgical steel inert Non-inflammatory, inert [1]
allows for ductile thin
vascular stent struts
Non-inflammatory,
Sturdy, highly suitable for
Titanium inert bone-friendly surface [1]
load-bearing applications
oxide layer

4. Polymeric Vascular Stents


Even though they may act somewhat inflammatory compared to metals, biodegradable polymeric
implants have been routinely employed as suture material and to temporarily fix tendons to bones until
they eventually adhere by themselves [37,38]. Popular hydrolysable polymers used for bioresorbable
scaffolds are poly(lactic-co-glycolic) acid (PLGA), polylactic acid (PLA) or polyglycolic acid (PGA) [39].
A main research focus has been polymeric stents, resulting in data that revealed several features that
had to be optimized. Since polymeric materials are generally less sturdy than metals, thicker struts

21
Metals 2018, 8, 532

are required. This results in stents that are more difficult to direct through small vessels. Moreover,
they are X-ray transparent and therefore harder to localize in the patient. Polymers also tend to swell
in aqueous environments and acidic hydrolysis products can act inflammatory [27]. In experimental
animal models degrading polymer stents resulted in increased restenosis rates [40,41]. One of the first
commercially available fully absorbable polylactic acid stent (Absorb, Abbot) that was FDA approved
in 2016 dissolved in two to three years, but despite promising short-term results long-term side
effects were negative and sales were terminated by 2017 [42,43]. In clinical trials these polymer stents
were more difficult to insert due to the increased efforts required for imaging, and over a two-year
period induced higher rates of in-stent thrombosis than drug eluting metal stents [44]. In summary,
the presently investigated resorbable polymer stents were deemed inferior to established metal stents.

5. Iron as a Prospective Stent Material


Pure iron and iron alloys were proposed in 2001 for corrodible stent materials [45]. Despite
appropriate mechanical properties, iron implants take years to disappear. The corrosion rate is an
order of magnitude too small for the implant to disappear without long-term side effects [30,46,47].
The immediate oxidation products (Fe2+ ) and ferrous (Fe3+ ) ions are essential for life and presumably
non-toxic at the expected concentrations [48–52]. In pioneering animal experiments iron implants
analysis revealed insoluble iron hydroxide precipitates that accumulated mainly at the site of
implantation [45,53]. Further analyses, in a mouse model, revealed iron precipitates engulfed by
local cells. After a few weeks these iron laden cells could be detected in various organs throughout
the body [54]. In war veterans, corroding iron fragments from grenade splinters have been shown to
migrate in the body and to cause chronic inflammation [55–57]. Overall, the slow degradation rate
prolonged possible side effects after completion of the healing process, and inflammatory precipitates
impede clinical applications of iron implants.

6. Zinc Alloy Stents


Corrodible zinc-based implants have been introduced relatively recently in 2013 (reviewed
in [58]). Even though the mechanical properties can be adjusted according to the requirements,
zinc alloys, with a reported yield strength up to 300 MPa, are not as strong as titanium or stainless
steel [59]. When tested, zinc alloys corroded with favorable kinetics, faster than iron, but less
rapidly than magnesium alloys. Zinc alloy degradation products were considered sufficiently
biocompatible [60]. In a rat model, zinc stents were still structurally intact after four months in
the abdominal aorta. The implant and the relevant degradation product Zn2+ appeared non-toxic
and even anti-inflammatory [61]. One year after the implantation of a pure zinc stent in a rabbit
aorta, an examination revealed artery remodeling and tissue healing without signs of inflammation,
platelet aggregation or thrombosis [33]. It was therefore concluded that selected zinc alloys had
promising strength and excellent biocompatibility for prospective bio-corrodible stent applications [62].
Nevertheless, it remains to be demonstrated in clinical trials that zinc alloys provide advantages over
clinically established permanent metal alloys.

7. Characteristics of Magnesium-Based Implants


The first reported medical application of degradable magnesium alloys in humans, as ligature
wire, was investigated in 1878 [63]. Side effects included the occurrence of gas pockets in the tissue,
and rapid, irregular pitting corrosion leading to premature implant failure. In part, pure magnesium
has been experimentally used to simplify the interpretation of biological responses. In general, alloy
metals such as aluminum, calcium, lithium, zirconium and rare earth elements have been used to
adjust mechanical properties such as the same stiffness as bony tissue or to reduce the degradation rate.
In addition, grain refinement, protective surface coatings, and metallic glasses obtained by ultrafast
cooling techniques resulted in improved degradation characteristics, increased material strength and
bone-compatible elastic moduli [64–75].

22
Metals 2018, 8, 532

In biological environments magnesium reacts with water molecules in a pitting type corrosion
with kinetics that depend on the surrounding tissue [76–78]. In addition, irregular corrosion could lead
to premature mechanical implant failure [79,80]. The primary magnesium corrosion products—soluble
magnesium ions (Mg2+ ), hydroxide ions (OH- ), and hydrogen gas (H2 )—are well tolerated by the
body. Mg2+ ions are essential for living cells, by complexing with the energy carrier adenosine
triphosphate and numerous enzymatic processes, and excess Mg2+ can be excreted in the urine [81,82].
Soluble hydroxide ions could in principle lead to toxic pH increases [76]. However, in biological
environments magnesium implants appear highly biocompatible presumably due to an adequate
buffering capacity of the tissue. In addition, magnesium and hydroxide ions combine in a pH neutral
way, and, together with carbonic acid, phosphates and other components present in surrounding
body fluids, precipitate to form a corrosion-retarding and highly biocompatible implant-tissue
interface [83,84]. However, perhaps initially surprisingly, during corrosion these precipitates can
transiently lead to increases of the overall implant mass and volume. This is particularly critical
for implants in non-yielding bony tissue. Stimulation of new bone growth and calcium phosphate
deposition has also been observed. This may be due to magnesium hydroxide deposition, calcium
phosphate precipitation at the tissue interface and the exertion of mechanical stress by the resulting
volume increase [85–87]. One gram of Mg can generate around one liter of hydrogen gas. Hydrogen
gas is non-toxic and easily diffusible, but excessive corrosion can nevertheless lead to formation
of undesirable gas bubbles (emphysema) in surrounding soft tissue. Excessive corrosion may also
lead to a buildup of pressure in bone enclosed cavities and may, therefore, stimulate bone growth in
appropriate setups [88,89].
In orthopedic applications selected magnesium alloys could achieve mechanical properties
more similar to human bone than titanium or steel. This could be favorable as it would reduce
implant-associated stress shielding and bone degradation [90,91]. Magnesium-based screws have
been used in bone healing clinical trials without notable side effects reported by patients [92,93].
The first commercial magnesium screws (Magnezix, Syntellix, Hannover, Germany) were available
in 2013, and completely disappeared one to two years after implantation [94]. Furthermore, recently
an additional interference screw, made of an MgYREZr-alloy, has been introduced to the market
(Milagro; DePuy Mitek, Leeds, United Kingdom) [95]. A transient appearance of radio translucent
areas around magnesium implants was reported [96]. In fact, such a phenomenon would be expected
from the above proposed mechanism; an initial magnesium implant size expansion by the deposition
and the subsequent resorption of solid corrosion products, leaving a temporary void space to be filled
by bony tissue.
Vascular magnesium alloy stents with reduced corrosion rates have been shown to be mechanically
stable for up to 6 months in animal experiments and were eventually evaluated in clinical trials [97–103].
Polymer-coated drug-eluting magnesium stents (Magmaris and DREAMS; Biotronik AG, 231, Bülach,
Switzerland) were commercially offered and claimed to be resorbed to 95% within a year in clinical
trials. Thus, they may thereby overcome long-term side effects [104–106]. Both orthopedic and vascular
magnesium implants appear promising but, with the exception of small orthopedic implants like pins
or screws, the development of these options is still in its infancy, and a broader clinical applicability
needs to be demonstrated [107].

8. Magnesium Implant Infection Susceptibility Mechanism: Race for the Surface versus
Susceptible Tissue Surface Model
Bacterial implant infections are difficult problem to treat in orthopedics, particularly in non-sterile
environments like the oral cavity [108]. Bacteria can form recalcitrant biofilms on implant surfaces that
are resistant to conventional antibiotic treatments. As a last resort, the entire implant may have to be
removed to allow an efficacious antibiotic treatment before the implant can be replaced. Corroding
magnesium has been shown to act as an antibacterial in vitro due to the generation of hydroxide
ions and pH increases [109–112]. In animal studies, an enhanced susceptibility to bacterial infections

23
Metals 2018, 8, 532

has been observed [113,114]. The reasons that could enhance the susceptibility to infection in vivo
are not understood, and difficult to explain. Any model must take into account that the corrosion
effects are no different in vitro, where there is no such enhanced susceptibility. The proposed model is
an attempt to explain this observation. Conventionally, exposed implant surfaces are thought to be
susceptible to bacterial adherence in competition with host tissue adhesion [115]. To allow bacterial
adhesion and survival on the freshly implanted magnesium, toxic pH increases directly at the interface
would have to be prevented in vivo. Unfortunately, experimental observation of the initial steps of
bacterial invasion has not been accomplished so far. However, this scenario appears unlikely if a
freshly implanted magnesium surface does act bactericidal. Importantly, despite systemic antibiotic
treatment, bacterial biofilms on magnesium were observed. Not only were they observed on the
implant surface but, also in the adjacent tissue (Figure 1), suggesting that bacterial adhesion to the
implant may actually not be essential for biofilm formation [113].
Alternatively, similar to burn wound infections or keratitis, initial bacterial invasion could occur
via the wound liquid to susceptible wounded tissue surfaces (Figure 2) [116,117]. If true for implanted
materials other than magnesium, this scenario would predict dire consequences for implant infection
prevention strategies.

Figure 1. Bacterial biofilm in tissue pockets at a distance from the implant surface. Magnesium
discs subcutaneously implanted into standard BALB/c mice were immediately infected with
Pseudomonas aeruginosa. After one week, tissue adjacent to the implants was subjected to scanning
transmission electron microscopic analysis (for a more detailed description see Reference [113]).
Bacteria (upper arrow) surrounded by clear areas (lower arrow), indicating the presence of
exopolysaccharide matrix material, a typical biofilm component. Reproduced with permission from
J. Biomed. Mater. Res.; Published by Wiley 2016.

24
Metals 2018, 8, 532

Figure 2. Model proposing tissue infection as initial key step of bacterial implant infections.
(A) Conventional model. Consecutive steps of biofilm infections are shown from left to right. Planktonic
bacteria (brown) enter the wound-liquid-filled interspace (colorless) between implant (grey) and tissue
(pink). As a crucial step towards biofilm formation, bacteria first adhere to the implant surface and
form micro-colonies. After reaching a critical density, bacteria switch to the biofilm mode and secrete
extracellular matrix compounds. Biofilm features, such as the encapsulation in the matrix, nutrient
restriction and slow growth, render the associated bacteria highly resistant to the host immune defenses
and to antibiotics. Subsequently, secreted exotoxins and proteases allow bacteria to invade the adjacent
host tissue. Alternatively, adhesion of host tissue to the implant acts to protect the implant surface
from bacterial attachment and subsequent biofilm formation. Based on the in vitro results, in this
scenario magnesium implants would be expected to act bactericidal. (B) Tissue infection model. Under
normal circumstances contiguous epithelial cell layers protect living tissue, whereas wounding renders
tissue highly susceptible to bacterial infections. After implant insertion the essential initial bacterial
attachment occurs primarily at the susceptible injured tissue surface. Bacterial colonies growing on the
tissue surface eventually switch to the biofilm mode with analogous outcomes as in the conventional
model. While bacterial adhesion to the implant may occur, it plays no essential role for the course of the
infection. Adhesion of host tissue to the implant would still be important to antagonize infections but
predominantly to protect the wound tissue surface, and not the implant, from bacterial colonization.
Despite acting bactericidal upon close contact, the observed enhanced infection susceptibility of
magnesium implants is explained by interference of corroding magnesium with host tissue adhesion.
Factors that prolong the wound surface exposure to bacteria could be alkaline pH immediately after
implantation, and hydrogen gas evolution or eroding solid corrosion layers thereafter.

9. Implications for the Design of Antibacterial Implants


A wide variety of anti-infective implant strategies have been investigated, mostly in vitro [118].
In the light of the proposed tissue invasion model, in order to be efficacious, antibacterial substances
would need to be diffusible to reach bacteria in the vicinity of the implant. Therefore, implant
nanostructures that act antiadhesive, or passive coatings that act bactericidal upon contact, would not
be expected to curb infections in patients. In addition, implant features that affect tissue adhesion play
an important, through different, role than previously thought. That is, to primarily prevent bacterial
adhesion to the injured tissue rather than to the implant (Table 2). Even though magnesium implants
could not curb bacterial infections in mice, clinical data is needed before a final conclusion can be
drawn. In addition, several alternative strategies are presently investigated, such as antibiotic-releasing
coatings for magnesium-based implants or the addition of antibacterial acting alloy metals like silver,
copper, or zinc that release cytotoxic ions [119–125]. The major challenge for such an approach is to
maintain the balance of achieving efficacious bactericidal ion concentrations in vivo without damaging
the host tissue.

25
Metals 2018, 8, 532

Table 2. Implant features predicted by the tissue infection model to influence the susceptibility to
infections in vivo.

Ineffective Coatings Infection Risks Favorable Measures


Factors that hinder host tissue Surfaces favoring tissue
Surfaces that antagonize
adhesion; convex or integration; smooth, flat or
bacterial adhesion
microporous surfaces concave forms;
Contact-dependent Relative movement of Antibacterial
bactericidal surfaces implants versus tissue substance-releasing coatings

10. Conclusions
In long-term clinical trials biodegradable polymeric stents were inferior to conventional
drug-eluting metal stents, while recently introduced biocorrodible magnesium-based bone screws
were without noticeable side effects. However, since in vitro tests and even small animal studies
cannot predict the outcome in human patients, long-term clinical confirmation of the expected benefits,
with regard to potential risks, are needed. In addition, a novel model for implant infections suggests
that host cell adhesion to implants is important to prevent bacterial invasion of the exposed host tissue
surface, and not, as previously thought, to prevent bacterial adhesion to the implant. The model
predicts that passive antibacterial implant coating strategies would not be efficacious in vivo.
Funding: We acknowledge the financial support by a joint grant to M.I.R. and to S.U. of the German
Academic Exchange Service (DAAD), Germany, and the Higher Education Commission of Pakistan
(HEC).

Conflicts of Interest: The authors declare that there is no conflict of interest regarding the publication of this paper.

References
1. Perren, S.M.; Regazzoni, P.; Fernandez, A.A. How to choose between the implant iaterials steel and titanium
in orthopedic trauma surgery: Part 2—Biological aspects. Acta Chir. Orthop. Traumatol. Cech. 2017, 84, 85–90.
[PubMed]
2. Osman, R.B.; Swain, M.V. A Critical Review of Dental Implant Materials with an Emphasis on Titanium
versus Zirconia. Materials (Basel) 2015, 8, 932–958. [CrossRef] [PubMed]
3. Williams, D.F. On the nature of biomaterials. Biomaterials 2009, 30, 5897–5909. [CrossRef] [PubMed]
4. Elias, C.N.; Lima, J.H.C.; Valiev, R.; Meyers, M.A. Biomedical applications of titanium and its alloys. JOM
2008, 60, 46–49. [CrossRef]
5. Sheikh, Z.; Najeeb, S.; Khurshid, Z.; Verma, V.; Rashid, H.; Glogauer, M. Biodegradable Materials for Bone
Repair and Tissue Engineering Applications. Materials (Basel) 2015, 8, 5744–5794. [CrossRef] [PubMed]
6. Middleton, J.C.; Tipton, A.J. Synthetic biodegradable polymers as orthopedic devices. Biomaterials 2000, 21,
2335–2346. [CrossRef]
7. Witte, F.; Fischer, J.; Nellesen, J.; Crostack, H.A.; Kaese, V.; Pisch, A.; Beckmann, F.; Windhagen, H. In vitro
and in vivo corrosion measurements of magnesium alloys. Biomaterials 2006, 27, 1013–1018. [CrossRef]
[PubMed]
8. Martinez Sanchez, A.H.; Luthringer, B.J.; Feyerabend, F.; Willumeit, R. Mg and Mg alloys: How comparable
are in vitro and in vivo corrosion rates? A review. Acta Biomater. 2015, 13, 16–31. [CrossRef] [PubMed]
9. Luthringer, B.J.; Feyerabend, F.; Willumeit-Romer, R. Magnesium-based implants: A mini-review.
Magnes. Res. 2014, 27, 142–154. [PubMed]
10. Li, H.; Zheng, Y.; Qin, L. Progress of biodegradable metals. Prog. Nat. Sci. 2014, 24, 414–422. [CrossRef]
11. Miller, A.P.; Huff, C.M.; Roubin, G.S. Vascular disease in the older adult. J. Geriatr. Cardiol. JGC 2016, 13,
727–732. [PubMed]
12. Martens, A.; Beckmann, E.; Kaufeld, T.; Umminger, J.; Fleissner, F.; Koigeldiyev, N.; Krueger, H.; Puntigam, J.;
Haverich, A.; Shrestha, M. Total aortic arch repair: Risk factor analysis and follow-up in 199 patients. Eur. J.
Cardiothorac. Surg. 2016, 50, 940–948. [CrossRef] [PubMed]

26
Metals 2018, 8, 532

13. Rittersma, S.Z.; de Winter, R.J.; Koch, K.T.; Bax, M.; Schotborgh, C.E.; Mulder, K.J.; Tijssen, J.G.; Piek, J.J.
Impact of strut thickness on late luminal loss after coronary artery stent placement. Am. J. Cardiol. 2004, 93,
477–480. [CrossRef] [PubMed]
14. Onche, I.I.; Osagie, O.E.; INuhu, S. Removal of orthopaedic implants: Indications, outcome and economic
implications. J. West Afr. Coll. Surg. 2011, 1, 101–112. [PubMed]
15. Levy, J.A.; Podeszwa, D.A.; Lebus, G.; Ho, C.A.; Wimberly, R.L. Acute complications associated with removal
of flexible intramedullary femoral rods placed for pediatric femoral shaft fractures. J. Pediatr. Orthop. 2013,
33, 43–47. [CrossRef] [PubMed]
16. Buchanan, K.; Steinvil, A.; Waksman, R. Does the new generation of drug-eluting stents render bare metal
stents obsolete? Cardiovasc. Revasc. Med. 2017, 18, 456–461. [CrossRef] [PubMed]
17. Waksman, R. A new generation of drug-eluting stents: Indications and outcomes of bioresorbable vascular
scaffolds. Cleve. Clin. J. Med. 2017, 84 (Suppl. S4), e20–e24. [CrossRef] [PubMed]
18. Pleva, L.; Kukla, P.; Hlinomaz, O. Treatment of coronary in-stent restenosis: A systematic review.
J. Geriatr. Cardiol. 2018, 15, 173–184. [PubMed]
19. Artang, R.; Dieter, R.S. Analysis of 36 reported cases of late thrombosis in drug-eluting stents placed in
coronary arteries. Am. J. Cardiol. 2007, 99, 1039–1043. [CrossRef] [PubMed]
20. Mauri, L.; Hsieh, W.H.; Massaro, J.M.; Ho, K.K.; D’Agostino, R.; Cutlip, D.E. Stent thrombosis in randomized
clinical trials of drug-eluting stents. N. Engl. J. Med. 2007, 356, 1020–1029. [CrossRef] [PubMed]
21. Filion, K.B.; Roy, A.M.; Baboushkin, T.; Rinfret, S.; Eisenberg, M.J. Cost-effectiveness of drug-eluting stents
including the economic impact of late stent thrombosis. Am. J. Cardiol. 2009, 103, 338–344. [CrossRef]
[PubMed]
22. Erne, P.; Schier, M.; Resink, T.J. The road to bioabsorbable stents: Reaching clinical reality? Cardiovasc.
Intervent. Radiol. 2006, 29, 11–16. [CrossRef] [PubMed]
23. Tan, L.; Yu, X.; Wan, P.; Yang, K. Biodegradable Materials for Bone Repairs: A Review. J. Mater. Sci. Technol.
2013, 29, 503–513. [CrossRef]
24. Zheng, Y.F.; Gu, X.N.; Witte, F. Biodegradable metals. Mater. Sci. Eng. R Rep. 2014, 77, 1–34. [CrossRef]
25. Waksman, R.; Pakala, R. Biodegradable and bioabsorbable stents. Curr. Pharm. Des. 2010, 16, 4041–4051.
[CrossRef] [PubMed]
26. Sumner, D.R. Long-term implant fixation and stress-shielding in total hip replacement. J. Biomech. 2015, 48,
797–800. [CrossRef] [PubMed]
27. Ceonzo, K.; Gaynor, A.; Shaffer, L.; Kojima, K.; Vacanti, C.A.; Stahl, G.L. Polyglycolic acid-induced
inflammation: Role of hydrolysis and resulting complement activation. Tissue Eng. 2006, 12, 301–308.
[CrossRef] [PubMed]
28. Athanasiou, K.A.; Niederauer, G.G.; Agrawal, C.M. Sterilization, toxicity, biocompatibility and clinical
applications of polylactic acid/polyglycolic acid copolymers. Biomaterials 1996, 17, 93–102. [CrossRef]
29. Pierson, D.; Edick, J.; Tauscher, A.; Pokorney, E.; Bowen, P.; Gelbaugh, J.; Stinson, J.; Getty, H.; Lee, C.H.;
Drelich, J.; et al. A simplified in vivo approach for evaluating the bioabsorbable behavior of candidate stent
materials. J. Biomed. Mater. Res. B Appl. Biomater. 2012, 100, 58–67. [CrossRef] [PubMed]
30. Bowen, P.K.; Drelich, J.; Buxbaum, R.E.; Rajachar, R.M.; Goldman, J. New approaches in evaluating metallic
candidates for bioabsorbable stents. Emerg. Mater. Res. 2012, 1, 237–255. [CrossRef]
31. Hermawan, H.; Purnama, A.; Dube, D.; Couet, J.; Mantovani, D. Fe-Mn alloys for metallic biodegradable
stents: Degradation and cell viability studies. Acta Biomater. 2010, 6, 1852–1860. [CrossRef] [PubMed]
32. Vojtech, D.; Kubasek, J.; Serak, J.; Novak, P. Mechanical and corrosion properties of newly developed
biodegradable Zn-based alloys for bone fixation. Acta Biomater. 2011, 7, 3515–3522. [CrossRef] [PubMed]
33. Yang, H.; Wang, C.; Liu, C.; Chen, H.; Wu, Y.; Han, J.; Jia, Z.; Lin, W.; Zhang, D.; Li, W.; et al. Evolution
of the degradation mechanism of pure zinc stent in the one-year study of rabbit abdominal aorta model.
Biomaterials 2017, 145, 92–105. [CrossRef] [PubMed]
34. Zhao, D.; Wang, T.; Nahan, K.; Guo, X.; Zhang, Z.; Dong, Z.; Chen, S.; Chou, D.T.; Hong, D.; Kumta, P.N.; et al.
In vivo characterization of magnesium alloy biodegradation using electrochemical H2 monitoring, ICP-MS,
and XPS. Acta Biomater. 2017, 50, 556–565. [CrossRef] [PubMed]
35. Zhao, D.; Witte, F.; Lu, F.; Wang, J.; Li, J.; Qin, L. Current status on clinical applications of magnesium-based
orthopaedic implants: A review from clinical translational perspective. Biomaterials 2017, 112, 287–302.
[CrossRef] [PubMed]

27
Metals 2018, 8, 532

36. Staiger, M.P.; Pietak, A.M.; Huadmai, J.; Dias, G. Magnesium and its alloys as orthopedic biomaterials:
A review. Biomaterials 2006, 27, 1728–1734. [CrossRef] [PubMed]
37. Debieux, P.; Franciozi, C.E.; Lenza, M.; Tamaoki, M.J.; Magnussen, R.A.; Faloppa, F.; Belloti, J.C. Bioabsorbable
versus metallic interference screws for graft fixation in anterior cruciate ligament reconstruction.
Cochrane Database Syst. Rev. 2016, 7, CD009772. [CrossRef] [PubMed]
38. Seitz, J.M.; Durisin, M.; Goldman, J.; Drelich, J.W. Recent Advances in Biodegradable Metals for Medical
Sutures: A Critical Review. Adv. Healthc. Mater. 2015, 4, 1915–1936. [CrossRef] [PubMed]
39. Doppalapudi, S.; Jain, A.; Khan, W.; Domb, A.J. Biodegradable polymers—An overview. Polym. Adv. Technol.
2014, 25, 427–435. [CrossRef]
40. Van der Giessen, W.J.; Lincoff, A.M.; Schwartz, R.S.; van Beusekom, H.M.; Serruys, P.W.; Holmes, D.R., Jr.;
Ellis, S.G.; Topol, E.J. Marked inflammatory sequelae to implantation of biodegradable and nonbiodegradable
polymers in porcine coronary arteries. Circulation 1996, 94, 1690–1697. [CrossRef] [PubMed]
41. Bunger, C.M.; Grabow, N.; Sternberg, K.; Kroger, C.; Ketner, L.; Schmitz, K.P.; Kreutzer, H.J.; Ince, H.;
Nienaber, C.A.; Klar, E.; et al. Sirolimus-eluting biodegradable poly-L-lactide stent for peripheral vascular
application: A preliminary study in porcine carotid arteries. J. Surg. Res. 2007, 139, 77–82. [CrossRef]
[PubMed]
42. Rizik, D.G.; Hermiller, J.B.; Kereiakes, D.J. Bioresorbable vascular scaffolds for the treatment of coronary
artery disease: Clinical outcomes from randomized controlled trials. Catheter. Cardiovasc. Interv. 2016, 88
(Suppl. S1), 21–30. [CrossRef] [PubMed]
43. Ali, Z.A.; Serruys, P.W.; Kimura, T.; Gao, R.; Ellis, S.G.; Kereiakes, D.J.; Onuma, Y.; Simonton, C.; Zhang, Z.;
Stone, G.W. 2-year outcomes with the Absorb bioresorbable scaffold for treatment of coronary artery disease:
A systematic review and meta-analysis of seven randomised trials with an individual patient data substudy.
Lancet 2017, 390, 760–772. [CrossRef]
44. Ali, Z.A.; Gao, R.; Kimura, T.; Onuma, Y.; Kereiakes, D.J.; Ellis, S.G.; Chevalier, B.; Vu, M.T.; Zhang, Z.;
Simonton, C.A.; et al. Three-Year Outcomes with the Absorb Bioresorbable Scaffold: Individual-Patient-Data
Meta-Analysis from the ABSORB Randomized Trials. Circulation 2018, 137, 464–479. [CrossRef] [PubMed]
45. Peuster, M.; Wohlsein, P.; Brugmann, M.; Ehlerding, M.; Seidler, K.; Fink, C.; Brauer, H.; Fischer, A.;
Hausdorf, G. A novel approach to temporary stenting: Degradable cardiovascular stents produced from
corrodible metal-results 6–18 months after implantation into New Zealand white rabbits. Heart 2001, 86,
563–569. [CrossRef] [PubMed]
46. Francis, A.; Yang, Y.; Virtanen, S.; Boccaccini, A.R. Iron and iron-based alloys for temporary cardiovascular
applications. J. Mater. Sci. Mater. Med. 2015, 26, 138. [CrossRef] [PubMed]
47. Peuster, M.; Hesse, C.; Schloo, T.; Fink, C.; Beerbaum, P.; von Schnakenburg, C. Long-term biocompatibility
of a corrodible peripheral iron stent in the porcine descending aorta. Biomaterials 2006, 27, 4955–4962.
[CrossRef] [PubMed]
48. Fagali, N.S.; Grillo, C.A.; Puntarulo, S.; Fernandez Lorenzo de Mele, M.A. Cytotoxicity of corrosion products
of degradable Fe-based stents: Relevance of pH and insoluble products. Colloids Surf. B Biointerfaces 2015,
128, 480–488. [CrossRef] [PubMed]
49. Moravej, M.; Purnama, A.; Fiset, M.; Couet, J.; Mantovani, D. Electroformed pure iron as a new biomaterial
for degradable stents: In vitro degradation and preliminary cell viability studies. Acta Biomater. 2010, 6,
1843–1851. [CrossRef] [PubMed]
50. Drynda, A.; Hoehn, R.; Peuster, M. Influence of Fe(II) and Fe(III) on the expression of genes related to
cholesterol- and fatty acid metabolism in human vascular smooth muscle cells. J. Mater. Sci. Mater. Med.
2010, 21, 1655–1663. [CrossRef] [PubMed]
51. Zhu, S.; Huang, N.; Xu, L.; Zhang, Y.; Liu, H.; Sun, H.; Leng, Y. Biocompatibility of pure iron: In vitro
assessment of degradation kinetics and cytotoxicity on endothelial cells. Mater. Sci. Eng. C 2009, 29,
1589–1592. [CrossRef]
52. Mueller, P.P.; May, T.; Perz, A.; Hauser, H.; Peuster, M. Control of smooth muscle cell proliferation by ferrous
iron. Biomaterials 2006, 27, 2193–2200. [CrossRef] [PubMed]
53. Kraus, T.; Moszner, F.; Fischerauer, S.; Fiedler, M.; Martinelli, E.; Eichler, J.; Witte, F.; Willbold, E.;
Schinhammer, M.; Meischel, M.; et al. Biodegradable Fe-based alloys for use in osteosynthesis: Outcome of
an in vivo study after 52 weeks. Acta Biomater. 2014, 10, 3346–3353. [CrossRef] [PubMed]

28
Metals 2018, 8, 532

54. Mueller, P.P.; Arnold, S.; Badar, M.; Bormann, D.; Bach, F.W.; Drynda, A.; Meyer-Lindenberg, A.; Hauser, H.;
Peuster, M. Histological and molecular evaluation of iron as degradable medical implant material in a
murine animal model. J. Biomed. Mater. Res. A 2012, 100, 2881–2889. [CrossRef] [PubMed]
55. Voigtlaender, H. [Wandering of foreign bodies. (Grenate fragment in the common bile duct causing jaundice)].
Chirurg 1975, 46, 467–469. [PubMed]
56. Ghislain, P.D. Spontaneous extrusion of hand grenade fragments from the face 60 years after injury. JAMA
2003, 290, 1317–1318. [CrossRef] [PubMed]
57. Gaber, Y. [Grenade splinter injury simulating thrombophlebitis]. Vasa 2003, 32, 40–42. [CrossRef] [PubMed]
58. Mostaed, E.; Sikora-Jasinska, M.; Drelich, J.W.; Vedani, M. Zinc-based alloys for degradable vascular stent
applications. Acta Biomater. 2018, 71, 1–23. [CrossRef] [PubMed]
59. Jin, H.; Zhao, S.; Guillory, R.; Bowen, P.K.; Yin, Z.; Griebel, A.; Schaffer, J.; Earley, E.J.; Goldman, J.; Drelich, J.W.
Novel high-strength, low-alloys Zn-Mg (<0.1 wt% Mg) and their arterial biodegradation. Mater. Sci. Eng. C
2018, 84, 67–79. [CrossRef] [PubMed]
60. Bowen, P.K.; Shearier, E.R.; Zhao, S.; Guillory, R.J., 2nd; Zhao, F.; Goldman, J.; Drelich, J.W. Biodegradable
Metals for Cardiovascular Stents: From Clinical Concerns to Recent Zn-Alloys. Adv. Healthc. Mater. 2016, 5,
1121–1140. [CrossRef] [PubMed]
61. Bowen, P.K.; Drelich, J.; Goldman, J. Zinc Exhibits Ideal Physiological Corrosion Behavior for Bioabsorbable
Stents. Adv. Mater. 2013, 25, 2577–2582. [CrossRef] [PubMed]
62. Bowen, P.K.; Guillory, R.J., 2nd; Shearier, E.R.; Seitz, J.M.; Drelich, J.; Bocks, M.; Zhao, F.; Goldman, J. Metallic
zinc exhibits optimal biocompatibility for bioabsorbable endovascular stents. Mater. Sci. Eng. C Mater.
Biol. Appl. 2015, 56, 467–472. [CrossRef] [PubMed]
63. Huse, E.C. A New Ligature? Chic. Med. J. Examin. 1878, 37, 171–172.
64. Agarwal, S.; Curtin, J.; Duffy, B.; Jaiswal, S. Biodegradable magnesium alloys for orthopaedic applications:
A review on corrosion, biocompatibility and surface modifications. Mater. Sci. Eng. C Mater. Biol. Appl. 2016,
68, 948–963. [CrossRef] [PubMed]
65. Waizy, H.; Seitz, J.-M.; Reifenrath, J.; Weizbauer, A.; Bach, F.-W.; Meyer-Lindenberg, A.; Denkena, B.;
Windhagen, H. Biodegradable magnesium implants for orthopedic applications. J. Mater. Sci. 2013, 48, 39–50.
[CrossRef]
66. Meagher, P.; O’Cearbhaill, E.D.; Byrne, J.H.; Browne, D.J. Bulk Metallic Glasses for Implantable Medical
Devices and Surgical Tools. Adv. Mater. 2016, 28, 5755–5762. [CrossRef] [PubMed]
67. Zberg, B.; Uggowitzer, P.J.; Loffler, J.F. MgZnCa glasses without clinically observable hydrogen evolution for
biodegradable implants. Nat. Mater. 2009, 8, 887–891. [CrossRef] [PubMed]
68. Frankel, G.S. Magnesium alloys: Ready for the road. Nat. Mater. 2015, 14, 1189–1190. [CrossRef] [PubMed]
69. Ma, E.; Xu, J. Biodegradable Alloys: The glass window of opportunities. Nat. Mater. 2009, 8, 855–857.
[CrossRef] [PubMed]
70. Li, Y.; Liu, L.; Wan, P.; Zhai, Z.; Mao, Z.; Ouyang, Z.; Yu, D.; Sun, Q.; Tan, L.; Ren, L.; et al. Biodegradable
Mg-Cu alloy implants with antibacterial activity for the treatment of osteomyelitis: In vitro and in vivo
evaluations. Biomaterials 2016, 106, 250–263. [CrossRef] [PubMed]
71. Xu, L.; Pan, F.; Yu, G.; Yang, L.; Zhang, E.; Yang, K. In vitro and in vivo evaluation of the surface bioactivity
of a calcium phosphate coated magnesium alloy. Biomaterials 2009, 30, 1512–1523. [CrossRef] [PubMed]
72. Wong, H.M.; Yeung, K.W.; Lam, K.O.; Tam, V.; Chu, P.K.; Luk, K.D.; Cheung, K.M. A biodegradable
polymer-based coating to control the performance of magnesium alloy orthopaedic implants. Biomaterials
2010, 31, 2084–2096. [CrossRef] [PubMed]
73. Li, X.; Chu, C.L.; Liu, L.; Liu, X.K.; Bai, J.; Guo, C.; Xue, F.; Lin, P.H.; Chu, P.K. Biodegradable poly-lactic acid
based-composite reinforced unidirectionally with high-strength magnesium alloy wires. Biomaterials 2015,
49, 135–144. [CrossRef] [PubMed]
74. Thormann, U.; Alt, V.; Heimann, L.; Gasquere, C.; Heiss, C.; Szalay, G.; Franke, J.; Schnettler, R.; Lips, K.S.
The Biocompatibility of Degradable Magnesium Interference Screws: An Experimental Study with Sheep.
BioMed Res. Int. 2015, 2015, 15. [CrossRef] [PubMed]
75. Hort, N.; Huang, Y.; Fechner, D.; Stormer, M.; Blawert, C.; Witte, F.; Vogt, C.; Drucker, H.; Willumeit, R.;
Kainer, K.U.; et al. Magnesium alloys as implant materials–principles of property design for Mg-RE alloys.
Acta Biomater. 2010, 6, 1714–1725. [CrossRef] [PubMed]

29
Metals 2018, 8, 532

76. Kirkland, N.T.; Birbilis, N.; Staiger, M.P. Assessing the corrosion of biodegradable magnesium implants:
A critical review of current methodologies and their limitations. Acta Biomater. 2012, 8, 925–936. [CrossRef]
[PubMed]
77. Huehnerschulte, T.A.; Angrisani, N.; Rittershaus, D.; Bormann, D.; Windhagen, H.; Meyer-Lindenberg, A.
In Vivo Corrosion of Two Novel Magnesium Alloys ZEK100 and AX30 and Their Mechanical Suitability as
Biodegradable Implants. Materials (Basel) 2011, 4, 1144–1167. [CrossRef] [PubMed]
78. Krause, A.; von der Höh, N.; Bormann, D.; Krause, C.; Bach, F.-W.; Windhagen, H.; Meyer-Lindenberg, A.
Degradation behaviour and mechanical properties of magnesium implants in rabbit tibiae. J. Mater. Sci.
2009, 45, 624. [CrossRef]
79. Zeng, R.; Dietzel, W.; Witte, F.; Hort, N.; Blawert, C. Progress and Challenge for Magnesium Alloys as
Biomaterials. Adv. Eng. Mater. 2008, 10, B3–B14. [CrossRef]
80. Kirkland, N.T.; Lespagnol, J.; Birbilis, N.; Staiger, M.P. A survey of bio-corrosion rates of magnesium alloys.
Corros. Sci. 2010, 52, 287–291. [CrossRef]
81. Jahnen-Dechent, W.; Ketteler, M. Magnesium basics. Clin. Kidney J. 2012, 5 (Suppl. S1), i3–i14. [CrossRef]
[PubMed]
82. De Baaij, J.H.; Hoenderop, J.G.; Bindels, R.J. Magnesium in man: Implications for health and disease.
Physiol. Rev. 2015, 95, 1–46. [CrossRef] [PubMed]
83. Han, P.; Cheng, P.; Zhang, S.; Zhao, C.; Ni, J.; Zhang, Y.; Zhong, W.; Hou, P.; Zhang, X.; Zheng, Y.; et al. In vitro
and in vivo studies on the degradation of high-purity Mg (99.99 wt.%) screw with femoral intracondylar
fractured rabbit model. Biomaterials 2015, 64, 57–69. [CrossRef] [PubMed]
84. Badar, M.; Lunsdorf, H.; Evertz, F.; Rahim, M.I.; Glasmacher, B.; Hauser, H.; Mueller, P.P. The formation of an
organic coat and the release of corrosion microparticles from metallic magnesium implants. Acta Biomater.
2013, 9, 7580–7589. [CrossRef] [PubMed]
85. Rahim, M.I.; Weizbauer, A.; Evertz, F.; Hoffmann, A.; Rohde, M.; Glasmacher, B.; Windhagen, H.; Gross, G.;
Seitz, J.M.; Mueller, P.P. Differential magnesium implant corrosion coat formation and contribution to bone
bonding. J. Biomed. Mater. Res. A 2017, 105, 697–709. [CrossRef] [PubMed]
86. Zhang, Y.; Xu, J.; Ruan, Y.C.; Yu, M.K.; O’Laughlin, M.; Wise, H.; Chen, D.; Tian, L.; Shi, D.; Wang, J.; et al.
Implant-derived magnesium induces local neuronal production of CGRP to improve bone-fracture healing
in rats. Nat. Med. 2016, 22, 1160–1169. [CrossRef] [PubMed]
87. Witte, F.; Kaese, V.; Haferkamp, H.; Switzer, E.; Meyer-Lindenberg, A.; Wirth, C.J.; Windhagen, H. In vivo
corrosion of four magnesium alloys and the associated bone response. Biomaterials 2005, 26, 3557–3563.
[CrossRef] [PubMed]
88. Kuhlmann, J.; Bartsch, I.; Willbold, E.; Schuchardt, S.; Holz, O.; Hort, N.; Hoche, D.; Heineman, W.R.; Witte, F.
Fast escape of hydrogen from gas cavities around corroding magnesium implants. Acta Biomater. 2013, 9,
8714–8721. [CrossRef] [PubMed]
89. Noviana, D.; Paramitha, D.; Ulum, M.F.; Hermawan, H. The effect of hydrogen gas evolution of magnesium
implant on the postimplantation mortality of rats. J. Orthop. Transl. 2016, 5, 9–15. [CrossRef]
90. Chen, Y.; Xu, Z.; Smith, C.; Sankar, J. Recent advances on the development of magnesium alloys for
biodegradable implants. Acta Biomater. 2014, 10, 4561–4573. [CrossRef] [PubMed]
91. Grünewald, T.A.; Rennhofer, H.; Hesse, B.; Burghammer, M.; Stanzl-Tschegg, S.E.; Cotte, M.; Löffler, J.F.;
Weinberg, A.M.; Lichtenegger, H.C. Magnesium from bioresorbable implants: Distribution and impact on
the nano- and mineral structure of bone. Biomaterials 2016, 76, 250–260. [CrossRef] [PubMed]
92. Seitz, J.-M.; Lucas, A.; Kirschner, M. Magnesium-Based Compression Screws: A Novelty in the Clinical Use
of Implants. JOM 2016, 68, 1177–1182. [CrossRef]
93. Plaass, C.; von Falck, C.; Ettinger, S.; Sonnow, L.; Calderone, F.; Weizbauer, A.; Reifenrath, J.; Claassen, L.;
Waizy, H.; Daniilidis, K.; et al. Bioabsorbable magnesium versus standard titanium compression screws for
fixation of distal metatarsal osteotomies—3 year results of a randomized clinical trial. J. Orthop. Sci. 2018, 23,
321–327. [CrossRef] [PubMed]
94. Windhagen, H.; Radtke, K.; Weizbauer, A.; Diekmann, J.; Noll, Y.; Kreimeyer, U.; Schavan, R.;
Stukenborg-Colsman, C.; Waizy, H. Biodegradable magnesium-based screw clinically equivalent to titanium
screw in hallux valgus surgery: Short term results of the first prospective, randomized, controlled clinical
pilot study. Biomed. Eng. Online 2013, 12, 62. [CrossRef] [PubMed]

30
Metals 2018, 8, 532

95. Ezechieli, M.; Meyer, H.; Lucas, A.; Helmecke, P.; Becher, C.; Calliess, T.; Windhagen, H.; Ettinger, M.
Biomechanical Properties of a Novel Biodegradable Magnesium-Based Interference Screw. Orthop. Rev.
(Pavia) 2016, 8, 6445. [CrossRef] [PubMed]
96. Biber, R.; Pauser, J.; Brem, M.; Bail, H.J. Bioabsorbable metal screws in traumatology: A promising innovation.
Trauma Case Rep. 2017, 8, 11–15. [CrossRef] [PubMed]
97. Ang, H.Y.; Huang, Y.Y.; Lim, S.T.; Wong, P.; Joner, M.; Foin, N. Mechanical behavior of polymer-based vs.
metallic-based bioresorbable stents. J. Thorac. Dis. 2017, 9 (Suppl. S9), S923–S934. [CrossRef] [PubMed]
98. Zhao, N.; Watson, N.; Xu, Z.G.; Chen, Y.J.; Waterman, J.; Sankar, J.; Zhu, D.H. In Vitro Biocompatibility and
Endothelialization of Novel Magnesium-Rare Earth Alloys for Improved Stent Applications. PLoS ONE
2014, 9, e98674. [CrossRef] [PubMed]
99. Zhu, S.J.; Liu, Q.; Qian, Y.F.; Sun, B.; Wang, L.G.; Wu, J.M.; Guan, S.K. Effect of different processings on
mechanical property and corrosion behavior in simulated body fluid of Mg-Zn-Y-Nd alloy for cardiovascular
stent application. Front. Mater. Sci. 2014, 8, 256–263. [CrossRef]
100. Mao, L.; Shen, L.; Chen, J.; Zhang, X.; Kwak, M.; Wu, Y.; Fan, R.; Zhang, L.; Pei, J.; Yuan, G.; et al. A promising
biodegradable magnesium alloy suitable for clinical vascular stent application. Sci. Rep. 2017, 7, 46343.
[CrossRef] [PubMed]
101. Mao, L.; Shen, L.; Chen, J.; Wu, Y.; Kwak, M.; Lu, Y.; Xue, Q.; Pei, J.; Zhang, L.; Yuan, G.; et al. Enhanced
bioactivity of Mg-Nd-Zn-Zr alloy achieved with nanoscale MgF2 surface for vascular stent application.
ACS Appl. Mater. Interfaces 2015, 7, 5320–5330. [CrossRef] [PubMed]
102. Waksman, R.; Erbel, R.; Di Mario, C.; Bartunek, J.; de Bruyne, B.; Eberli, F.R.; Erne, P.; Haude, M.; Horrigan, M.;
Ilsley, C.; et al. Early- and long-term intravascular ultrasound and angiographic findings after bioabsorbable
magnesium stent implantation in human coronary arteries. JACC Cardiovasc. Interv. 2009, 2, 312–320.
[CrossRef] [PubMed]
103. Erbel, R.; Di Mario, C.; Bartunek, J.; Bonnier, J.; de Bruyne, B.; Eberli, F.R.; Erne, P.; Haude, M.; Heublein, B.;
Horrigan, M.; et al. Temporary scaffolding of coronary arteries with bioabsorbable magnesium stents:
A prospective, non-randomised multicentre trial. Lancet 2007, 369, 1869–1875. [CrossRef]
104. Rapetto, C.; Leoncini, M. Magmaris: A new generation metallic sirolimus-eluting fully bioresorbable scaffold:
Present status and future perspectives. J. Thorac. Dis. 2017, 9 (Suppl. S9), S903–S913. [CrossRef] [PubMed]
105. Haude, M.; Ince, H.; Kische, S.; Abizaid, A.; Tolg, R.; Alves Lemos, P.; Van Mieghem, N.M.; Verheye, S.;
von Birgelen, C.; Christiansen, E.H.; et al. Sustained safety and clinical performance of a drug-eluting
absorbable metal scaffold up to 24 months: Pooled outcomes of BIOSOLVE-II and BIOSOLVE-III.
EuroIntervention 2017, 13, 432–439. [PubMed]
106. Haude, M.; Erbel, R.; Erne, P.; Verheye, S.; Degen, H.; Bose, D.; Vermeersch, P.; Wijnbergen, I.; Weissman, N.;
Prati, F.; et al. Safety and performance of the drug-eluting absorbable metal scaffold (DREAMS) in patients
with de-novo coronary lesions: 12 month results of the prospective, multicentre, first-in-man BIOSOLVE-I
trial. Lancet 2013, 381, 836–844. [CrossRef]
107. Willbold, E.; Weizbauer, A.; Loos, A.; Seitz, J.M.; Angrisani, N.; Windhagen, H.; Reifenrath, J. Magnesium
alloys: A stony pathway from intensive research to clinical reality. Different test methods and
approval-related considerations. J. Biomed. Mater. Res. A 2017, 105, 329–347. [CrossRef] [PubMed]
108. Arciola, C.R.; Campoccia, D.; Speziale, P.; Montanaro, L.; Costerton, J.W. Biofilm formation in Staphylococcus
implant infections. A review of molecular mechanisms and implications for biofilm-resistant materials.
Biomaterials 2012, 33, 5967–5982. [CrossRef] [PubMed]
109. Rahim, M.I.; Eifler, R.; Rais, B.; Mueller, P.P. Alkalization is responsible for antibacterial effects of corroding
magnesium. J. Biomed. Mater. Res. A 2015, 103, 3526–3532. [CrossRef] [PubMed]
110. Feng, H.; Wang, G.; Jin, W.; Zhang, X.; Huang, Y.; Gao, A.; Wu, H.; Wu, G.; Chu, P.K. Systematic Study of
Inherent Antibacterial Properties of Magnesium-based Biomaterials. ACS Appl. Mater. Interfaces 2016, 8,
9662–9673. [CrossRef] [PubMed]
111. Lock, J.Y.; Wyatt, E.; Upadhyayula, S.; Whall, A.; Nunez, V.; Vullev, V.I.; Liu, H. Degradation and antibacterial
properties of magnesium alloys in artificial urine for potential resorbable ureteral stent applications. J. Biomed.
Mater. Res. A 2014, 102, 781–792. [CrossRef] [PubMed]
112. Robinson, D.A.; Griffith, R.W.; Shechtman, D.; Evans, R.B.; Conzemius, M.G. In vitro antibacterial
properties of magnesium metal against Escherichia coli, Pseudomonas aeruginosa and Staphylococcus aureus.
Acta Biomater. 2010, 6, 1869–1877. [CrossRef] [PubMed]

31
Metals 2018, 8, 532

113. Rahim, M.I.; Rohde, M.; Rais, B.; Seitz, J.-M.; Mueller, P.P. Susceptibility of metallic magnesium implants to
bacterial biofilm infections. J. Biomed. Mater. Res. A 2016, 104, 1489–1499. [CrossRef] [PubMed]
114. Hou, P.; Zhao, C.; Cheng, P.; Wu, H.; Ni, J.; Zhang, S.; Lou, T.; Wang, C.; Han, P.; Zhang, X.; et al. Reduced
antibacterial property of metallic magnesium in vivo. Biomed. Mater. 2016, 12, 015010. [CrossRef] [PubMed]
115. Gristina, A.G.; Naylor, P.; Myrvik, Q. Infections from biomaterials and implants: A race for the surface.
Med. Prog. Technol. 1988, 14, 205–224. [PubMed]
116. Church, D.; Elsayed, S.; Reid, O.; Winston, B.; Lindsay, R. Burn wound infections. Clin. Microbiol. Rev. 2006,
19, 403–434. [CrossRef] [PubMed]
117. Stern, G.A. Pseudomonas keratitis and contact lens wear: The lens/eye is at fault. Cornea 1990, 9 (Suppl. S1),
S39–S40. [CrossRef] [PubMed]
118. Qin, S.; Xu, K.; Nie, B.; Ji, F.; Zhang, H. Approaches based on passive and active antibacterial coating on
titanium to achieve antibacterial activity. J. Biomed. Mater. Res. A 2018. [CrossRef] [PubMed]
119. Liu, Z.D.; Schade, R.; Luthringer, B.; Hort, N.; Rothe, H.; Muller, S.; Liefeith, K.; Willumeit-Romer, R.;
Feyerabend, F. Influence of the Microstructure and Silver Content on Degradation, Cytocompatibility, and
Antibacterial Properties of Magnesium-Silver Alloys In Vitro. Oxid. Med. Cell. Longev. 2017, 2017, 8091265.
[CrossRef] [PubMed]
120. Zhang, X.D.; Yi, J.H.; Zhao, G.W.; Huang, L.L.; Yan, G.J.; Chen, Y.S.; Liu, P. Layer-by-layer assembly of
silver nanoparticles embedded polyelectrolyte multilayer on magnesium alloy with enhanced antibacterial
property. Surf. Coat. Technol. 2016, 286, 103–112. [CrossRef]
121. Yu, W.L.; Chen, D.Y.; Ding, Z.Y.; Qiu, M.L.; Zhang, Z.W.; Shen, J.; Zhang, X.N.; Zhang, S.X.; He, Y.H.; Shi, Z.M.
Synergistic effect of a biodegradable Mg-Zn alloy on osteogenic activity and anti-biofilm ability: An in vitro
and in vivo study. RSC Adv. 2016, 6, 45219–45230. [CrossRef]
122. Zhao, C.; Hou, P.; Ni, J.; Han, P.; Chai, Y.; Zhang, X. Ag-Incorporated FHA Coating on Pure Mg: Degradation
and in Vitro Antibacterial Properties. ACS Appl. Mater. Interfaces 2016, 8, 5093–5103. [CrossRef] [PubMed]
123. Qin, H.; Zhao, Y.; An, Z.; Cheng, M.; Wang, Q.; Cheng, T.; Wang, Q.; Wang, J.; Jiang, Y.; Zhang, X.; et al.
Enhanced antibacterial properties, biocompatibility, and corrosion resistance of degradable Mg-Nd-Zn-Zr
alloy. Biomaterials 2015, 53, 211–220. [CrossRef] [PubMed]
124. Tie, D.; Feyerabend, F.; Hort, N.; Hoeche, D.; Kainer, K.U.; Willumeit, R.; Mueller, W.D. In vitro mechanical
and corrosion properties of biodegradable Mg-Ag alloys. Mater. Corros. 2014, 65, 569–576. [CrossRef]
125. Zhang, X.B.; Ba, Z.X.; Wang, Z.Z.; He, X.C.; Shen, C.; Wang, Q. Influence of silver addition on microstructure
and corrosion behavior of Mg-Nd-Zn-Zr alloys for biomedical application. Mater. Lett. 2013, 100, 188–191.
[CrossRef]

© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

32
metals
Article
Influence of the Composition of the Hank’s Balanced
Salt Solution on the Corrosion Behavior of AZ31 and
AZ61 Magnesium Alloys
Jakub Tkacz 1,2, *, Karolína Slouková 1 , Jozef Minda 1 , Juliána Drábiková 1 , Stanislava Fintová 1,3 ,
Pavel Doležal 1,4 and Jaromír Wasserbauer 1
1 Materials Research Centre, Faculty of Chemistry, Brno University of Technology, 612 00 Brno,
Czech Republic; xcsloukova@fch.vut.cz (K.S.); xcminda@fch.vut.cz (J.M.); xcdrabikovaj@fch.vut.cz (J.D.);
fintova@ipm.cz (S.F.); dolezal@fme.vutbr.cz (P.D.); wasserbauer@fch.vut.cz (J.W.)
2 Research Centre, University of Žilina, 010 08 Žilina, Slovakia
3 Institute of Physics of Materials AS CR v. v. i., Žižkova 22, 616 62 Brno, Czech Republic
4 Institute of Materials Science and Engineering, Faculty of Mechanical Engineering,
Brno University of Technology, Technická 2896/2, 616 69 Brno, Czech Republic
* Correspondence: tkacz@fch.vut.cz; Tel.: +420-54-114-9469

Received: 31 August 2017; Accepted: 25 October 2017; Published: 1 November 2017

Abstract: The electrochemical corrosion characteristics of AZ31 and AZ61 magnesium alloys
were analyzed in terms of potentiodynamic tests and electrochemical impedance spectroscopy.
The influence of the solution composition and material surface finish was examined also through the
analysis of corrosion products created on the samples’ surface after electrochemical measurements
in terms of scanning electron microscopy using energy-dispersive spectroscopy. Obtained data
revealed the differences in the response of the magnesium alloys to enriched Hank’s Balanced
Salt Solution—HBSS+ (with Mg2+ and Ca2+ ions) and Hank’s Balanced Salt Solution—HBSS
(without Mg2+ and Ca2+ ions). Both examined alloys exhibited better corrosion resistance from
the thermodynamic and kinetic point of view in the enriched HBSS+. AZ61 magnesium alloy reached
higher values of polarization resistance than AZ31 magnesium alloy in both the used corrosion
solutions. Phosphate-based corrosion products were characteristic for the AZ31 and AZ61 alloys
tested in the HBSS (without Mg2+ and Ca2+ ions). The combination of phosphate-based corrosion
products and clusters of MgO and Mg(OH)2 was typical for the surface of samples tested in the
enriched HBSS+ (with Mg2+ and Ca2+ ions). Pitting corrosion attack was observed only in the case of
enriched HBSS+.

Keywords: magnesium alloy; AZ31; AZ61; HBSS; HBSS+; EIS; potentiodynamic test

1. Introduction
Magnesium is an essential element for living organisms, however, for technical purposes
is the magnesium used mainly in the form of alloys. Alloying elements improve magnesium
mechanical properties and it can be used to control its reactivity. Due to the suitable combination
of physico-mechanical properties, biocompatibility and non-toxicity specific magnesium alloys are
investigated for medical applications. In the case of orthopedic implants physical and mechanical
properties of magnesium alloys are also important. These properties are similar to the properties of a
human bone (e.g., density, compressive yield strength, ultimate tensile strength). Magnesium alloy
implants are moreover biocompatible and biodegradable [1–11]. As a result of chemical reactions with
the biological environment non-toxic corrosion products are created on the surface of the implants.
In the human body the magnesium alloy implants dissolve and are absorbed, which prevents surgical

Metals 2017, 7, 465; doi:10.3390/met7110465 33 www.mdpi.com/journal/metals


Metals 2017, 7, 465

removal of the implants after tissue healing [5,6]. The disadvantage of magnesium and magnesium
alloys is their high reactivity at the physiological pH (7.4–7.6) as well as in physiological media
containing high concentrations of chloride ions, which could cause rapid disintegration of the implant
in the biological environment [7,8]. Furthermore, during the corrosion process of magnesium and its
alloys, the release of hydrogen gas may be too fast to be endured by the host tissues [9].
One way to influence corrosion resistance and mechanical properties of magnesium alloys is by
using high purity alloys that maintain metal impurities such as iron, nickel and copper below limits.
Examples of alloying elements of magnesium alloys for biodegradable implants for improvement of the
corrosion resistance of the alloys are calcium, zinc, etc. [10–15]. On the other hand, even the magnesium
alloys for medical applications have to have good mechanical properties and they have to also contain
other alloying elements. One of the basic alloying elements for magnesium mechanical and corrosion
properties improvement is aluminum. Even though Al has a positive effect on magnesium alloys
properties, the amount of Al added must be controlled in the case of alloys for medical applications.
A high concentration of Al was considered to possibly cause neurotoxic illnesses such as dementia or
Alzheimer’s disease. However, when Al is introduced into the human body in small concentrations,
for instance during dietary ingestion or consumption from natural or urban water supplies, then it is
naturally excreted through urine or in the form of bile [10,16–21].
Hank’s balanced salt solutions (HBSSs), which are one of the options for simulating the corrosive
environment of the body of living organisms are often used for analysis of the corrosion behavior
of magnesium alloys that are expected to be used in medicine applications. Mainly due to higher
chloride concentrations, HBSSs are more aggressive medium compared to artificial plasma. Sulfate ions
contained in HBSS can also result in higher corrosion rate of magnesium and its alloys compared to
other corrosion media used for material corrosion properties characterization [22–31].
The reactivity of material in the corrosion environment is, besides many aspects, influenced with
the chemical and phase composition of the material and chemical composition of the testing solution.
Electrochemical corrosion behavior of wrought AZ31 and AZ61 alloys in HBSS characterized by
Tkacz et al. in [32] revealed different response of the materials due to their chemical and phase
composition and surface finish. Based on the EIS (electrochemical impedance spectroscopy)
measurement results, AZ61 magnesium alloy was considered as more corrosion resistant when
compared to AZ31 magnesium alloy, while opposite conclusion can be considered based on the
potentiodynamic test results. Potentiodynamic tests revealed minor influence of the surface finish in
the case of AZ31 magnesium alloy represented for example by corrosion potential (Ecorr ) values.
The samples with polished surface (Ecorr = −1.676 ± 0.003 V) were characteristic with more
positive value of Ecorr comparing to the ground samples (Ecorr = −1.701 ± 0.003 V). On the
other hand, no influence of surface finish was observed in the case of AZ61 magnesium alloy
(Ecorr = −1.708 ± 0.004 V for ground sample and Ecorr = −1.708 ± 0.003 V for polished sample).
No significant influence of surface finish was observed based on the EIS test results. The increase
of polarization resistance up to 24 h of immersion of the samples in the HBSS to the values of
approximately 4000 Ω·cm2 was characteristic for both the surface states of AZ31 magnesium alloy,
while increasing exposure time to the corrosion environment did not have any significant influence
on the polarization resistance and the value remained stable. In the case of AZ61 magnesium alloy a
significant increase of the values of polarization resistance up to the 48 h of exposure of the samples to
the corrosion environment was observed for both the material states, reaching the value of polarization
resistance of approximately 9000 Ω·cm2 . Following an increase of immersion time resulted in an
additional increase of the Rp to the value of approximately 21,000 Ω·cm2 for ground and 17,000 Ω·cm2
for polished samples.
The influence of the chloride ions on corrosion behavior of Mg-Al-Zn based alloys was reported
in several studies. Ambat et al. studied in [33] the influence of chloride ion concentration and pH
on the corrosion and electrochemical behavior of die-cast and ingot-cast AZ91D alloy. The effect of
chloride ion concentration was studied in NaCl (0–10%) solutions at pH 7.25. The effect of pH was

34
Metals 2017, 7, 465

analyzed in solutions with the chloride ion concentration kept constant at 3.5% while the pH was
varied from 1.0–12.0. Material behavior was analyzed with immersion and potentiodynamic testing.
Increase in chloride ion concentration increased the corrosion rate of AZ91D magnesium alloy at
pH 7.25 and 12.0 for both the material states, however, at pH 2.0, the effect of chloride ion was found to
be negligible. High corrosion rate was observed for both the material states in highly acidic solutions,
while the corrosion rate was found to be low in neutral pH and alkaline conditions. The corrosion
rate determined from immersion tests was much higher than that obtained from electrochemical
measurements. The observation was attributed to the negative difference effect and to the physical
removal of β phase during corrosion. The differences between the response of material states to the
corrosion environment were discussed in terms of microstructural differences.
The effect of chloride ion concentration and pH on the corrosion (immersion tests) and
electrochemical behavior (potentiodynamic tests) of AZ63 alloy were studied in NaCl solutions at
different concentrations (0.01, 0.2, 0.6, 1 and 2 M) and pH values (2, 3, 8, 11 and 11.5) were studied
by Altun and Sen in [34]. Authors observed that the corrosion rate increased with the increase in
concentration of NaCl solution. However, it was observed, that with the increase in chloride ion
concentration, the rising rate of corrosion rate decreased. The corrosion rate increase with increasing
chloride ion concentration was attributed to the participation of chloride ions in the dissolution
reaction. Chloride ions are aggressive for both magnesium and aluminum. The adsorption of chloride
ions to oxide covered magnesium surface transforms Mg(OH)2 to easily soluble MgCl2 . Authors also
observed a shift of the corrosion potential to more negative (more active) values with the increase in
chloride ion concentration. The explanation for this behavior was found in adsorption of these ions on
the alloy surface at weak parts of the oxide film. Corrosion potential was observed to be shifted to
more negative (more active) values with the decrease in pH value of the solution. Higher pH values
were discussed to favor the formation of Mg(OH)2 which protects the alloy from corrosion.
Merino et al. studied in [35] the influence of chloride ion concentration and temperature on the
corrosion of Mg-Al alloys in a salt fog with the focus on the effect of Al content in the alloy. The results
of their investigation showed that the corrosion attack of Mg, AZ31, AZ80 and AZ91D materials under
the salt fog test increases with increasing temperature and chloride anion concentration, while the effect
of temperature was considered to be more noticeable than that of chloride concentration. Authors also
analyzed the influence of the Al content and resulting presence of intermetallic phases on material
corrosion behavior. In the case of the wrought AZ31 only negligible influence of the present AlMn
based phase was observed, while in the case of cast AZ80 and AZ91 alloys the creation of galvanic
couples between Al-Mn and β-Mg17 Al12 phases with the Mg matrix resulted in more pronounced
corrosion attack. Authors also observed the influence of the distribution, size and morphology of
the β phase and resulting Al-rich corrosion products layer created on the material surface during
the corrosion.
Influence of sulfate anion concentration and pH on the corrosion of Mg-Al-Zn-Mn (GA9)
magnesium alloy was investigated by Shetty et al. [36]. The studies were carried out in sodium
sulfate solutions with concentrations range of 0.1–2 M; and at different temperatures of 30–50 ◦ C and
pH of 3.0–12.0. According to the experimental data, the corrosion rate of the alloy increased with
the increase in temperature, and also with the increase in the concentration of sodium sulfate in the
medium. It was observed that the rate of corrosion decreased with the increase in pH.
Even thought, magnesium alloys corrosion properties are widely investigated, most of the studies
are performed in NaCl and Na2 SO4 solutions simulating corrosion environment in engineering
applications [33–36]. Corrosion behavior of magnesium alloys was studied in several types of
HBSS [22–32], however, the influence of the chemical composition of HBSS on corrosion processes is
not available in the literature according to the authors’ knowledge.
Corrosion characteristics of metallic materials can be analyzed in different ways. In this work,
electrochemical methods have been used to investigate the corrosion behavior of magnesium alloys.
Potentiodynamic tests and electrochemical impedance spectroscopy were used for the description of

35
Metals 2017, 7, 465

the material response to the HBSS and enriched HBSS+. Obtained data were discussed with the aim to
identify the influence of material chemical and phase composition, surface roughness and composition
of the corrosion solution on corrosion behavior. Values of corrosion potential (Ecorr ) expressing
thermodynamics of the corrosion process and corrosion current density values (icorr ) expressing the
kinetic of the corrosion process were obtained by potentiodynamic measurements. Corrosion potential
characteristic for the material expresses the thermodynamic stability of the system and the conditions
for material corrosion and its resistance against corrosion process. Kinetic of the corrosion process
can be shown by the evolution of the corrosion rate (vcorr ) which can be calculated from the icorr .
Polarization resistance (Rp ) values were obtained by electrochemical impedance spectroscopy (EIS).
Evolution of the corrosion products on the specimen surface was analyzed in terms of scanning electron
microscopy and correlated with the composition of the used corrosion solution.
The presented results show differences in electrochemical corrosion behavior of AZ31 and
AZ61 alloys in HBSS and enriched HBSS+ with the aim to characterize material behavior and
different response of ground and polished materials on the different chemical composition of the
corrosion solution.

2. Materials and Methods

2.1. Material
Wrought AZ31 and AZ61 magnesium alloys plates were used for the experiments. Metallographic
analysis and verification of chemical composition of the magnesium alloys were performed by
scanning electron microscope (SEM) (ZEISS EVO LS 10, Cambridge, UK) with energy dispersive
spectrometer (EDS) (OXFORD Instruments X-MAX 80 mm2 , Abingdon, UK). Metallographic samples
for microstructural analysis were prepared in terms of a conventional procedure consisting of grinding,
polishing (diamond paste 1 μm) and etching (picral solution [37]).

2.2. Electrochemical Measurements


Wrought AZ31 and AZ61 alloys plates were cut to samples with dimensions of 20 × 20 × 2 mm3 .
One batch of the samples of each magnesium alloy was ground with 1200 grit SiC paper with a particle
size of ~15 μm (marked #1200) and the second batch was additionally polished with diamond pastes
up to 0.25 μm (marked 0.25 μm). Wetting agent during polishing was isopropyl alcohol.
Electrochemical characteristics of the ground and polished samples were measured by
potentiostat/galvanostat BioLogic VSP-300 (BioLogic, Seyssinet-Pariset, France). Three-electrode
system was used for electrochemical measurements, where the magnesium alloy sample served as
a working electrode (WE), saturated calomel electrode (SCE) was used as a reference electrode (RE)
and platinum gauze was used as a counter electrode (CE). Used corrosion environment was enriched
HBSS+ (with the addition of Mg2+ and Ca2+ ions) and HBSS (without Mg2+ and Ca2+ ions) from [32]
was used to show the dependence of the response of the magnesium alloys on corrosion solution
composition. Experiments were performed at the temperature of 37 ± 1 ◦ C. The composition of the
used HBSS and enriched HBSS+ is given in Table 1. An area of 1.0 cm2 of the sample was exposed to
the corrosion environment during the electrochemical testing. The stabilization time of the sample
exposed to the corrosion environment before the measurement was 5 min. Each electrochemical
measurement was performed on three specimens with adequate surface finish.

Table 1. Chemical composition of HBSS and enriched HBSS+ used in [32].

Composition (mg·dm−3 )
Solutions
NaCl KCl KH2 PO4 Glucose Na2 HPO4 MgSO4 CaCl2 Na2 CO3
HBSS 8000 400 60 1000 48 - - 350
Enriched HBSS+ 8000 400 60 1000 48 98 140 350

36
Metals 2017, 7, 465

Potentiodynamic measurements were performed by polarizing of the sample surface in the range
from −100 mV to +200 mV vs. open circuit potential (EOCP ). Scan rate was 1 mV·s−1 .
Electrochemical impedance spectroscopy (EIS) measurements were performed after exposure of
the sample surface to the enriched HBSS+ with Mg2+ and Ca2+ ions for 5 min, 1, 2, 4, 8, 12, 24, 48, 72,
96 and 168 h. EIS scan frequency ranged from 100 kHz to 100 mHz, and the perturbation amplitude
was 5 mV.
With the aim to determine the influence of the enriched HBSS+ with Mg2+ and Ca2+ ions on the
experimental material electrochemical corrosion behavior were the EIS measurements performed also
in the solution not containing addition Mg2+ and Ca2+ ions (Table 1).
The chemical composition of the corrosion products created on the surface of the AZ31 and
AZ61 alloys during EIS measurements was analyzed with SEM ZEISS EVO LS 10 with EDS OXFORD
Instruments X-MAX 80 mm2 . The samples were rinsed with isopropyl alcohol and dried with air
before the analysis.

3. Results

3.1. Microstructural Analysis


The microstructure of the AZ31 magnesium alloy is consisted of polyhedral grains of the
substitutional solid solution α-Mg in which AlMn intermetallic phase particles were observed
(Figure 1) [38–40]. The distribution of the basic alloying elements revealed with the mapping mode by
EDS is shown in Figure 1. The chemical composition of the AZ31 magnesium alloy verified by EDS
is provided in Table 2. The content of the main alloying elements (Al, Zn and Mn) agrees with the
standard ASTM B90M [41].

Figure 1. Microstructure of AZ31 magnesium alloy SEM (Scanning Electron Microscope) with EDS
(Energy Dispersive Spectrometer) maps of the elements distribution: magnesium (Mg), aluminum (Al),
zinc (Zn) and manganese (Mn).

Table 2. Chemical composition of AZ31 magnesium alloy (EDS: Energy Dispersive Spectrometer).

Element Al Zn Mn Mg
Chemical composition (wt %) 3.2 0.9 0.4 balance

The microstructure of the AZ61 magnesium alloy is consists of polyhedral grains of the
substitution solid solution (α-Mg), β-phase (particles of Mg17 Al12 intermetallic phase) and AlMn
intermetallic phase particles (Figure 2). The EDS mapping mode was used for the investigation of the
distribution of the basic alloying elements in the alloy (Figure 2). The chemical composition of the

37
Metals 2017, 7, 465

AZ61 magnesium alloy was verified by EDS (Table 3). The content of the main alloying elements (Al,
Zn and Mn) agrees with the standard ASTM B107M [42].

Figure 2. Microstructure of AZ61 magnesium alloy (SEM) with EDS maps of the elements distribution:
magnesium (Mg), aluminum (Al), zinc (Zn) and manganese (Mn).

Table 3. Chemical composition of AZ61 magnesium alloy (EDS).

Element Al Zn Mn Mg
Chemical composition (wt %) 6.0 0.7 0.3 balance

3.2. Potentiodynamic Measurements


Figure 3 shows typical potentiodynamic curves obtained with linear polarization of ground and
polished samples of AZ31 and AZ61 alloys. The pitting corrosion attack of all the tested samples
was revealed by an increase of current density at the anodic branch of the curves which belongs to
the value of pitting potential (Epit ) [38]. The values of the corrosion current density (icorr ) for the
AZ31 magnesium alloy were determined only from the cathodic branch of the polarization curves
applying the Tafel extrapolation. The Tafel region for the anodic branches of the potentiodynamic
curves characterizing AZ31 magnesium alloy was insufficient to use the Tafel extrapolation to obtain
relevant values of icorr , following the rule that the plane region has to be at least 50 mV [38] from the
Ecorr . On the other hand, icorr values for the AZ61 magnesium alloy were determined from both the
branches of the obtained polarization curves applying Tafel extrapolation.

Figure 3. Potentiodynamic curves of ground (#1200) and polished (0.25 μm) surface of AZ31 and AZ61
magnesium alloys in enriched HBSS+ (with Mg2+ and Ca2+ ions).

38
Metals 2017, 7, 465

For both the tested materials the finer surface (polishing with 0.25 μm diamond paste) resulted
in the potentiodynamic curve shift to more positive values of potential comparing to the curves
characterizing the ground samples.
The electrochemical characteristics estimated by potentiodynamic measurements: EOCP , Ecorr ,
Epit , icorr and corrosion rate vcorr are given in Table 4. From the obtained data, a larger influence of
the surface finish can be observed in the case of AZ61 magnesium alloy when compared to the AZ31
magnesium alloy.

Table 4. Results of potentiodynamic tests performed in enriched HBSS+ (with Mg2+ and Ca2+ ions).

Alloy Finish EOCP (V) Ecorr (V) Epit (V) icorr (μA·cm−2 ) vcorr (μm·year−1 )
AZ31—#1200 −1.721 ± 0.015 −1.556 ± 0.012 −1.500 ± 0.014 10.5 ± 0.8 234.6 ± 56.6
AZ61—#1200 −1.717 ± 0.014 −1.594 ± 0.014 −1.509 ± 0.009 5.5 ± 0.6 123.9 ± 40.6
AZ31—0.25 μm −1.674 ± 0.005 −1.538 ± 0.008 −1.474 ± 0.011 4.9 ± 0.5 112.5 ± 26.8
AZ61—0.25 μm −1.643 ± 0.013 −1.505 ± 0.011 −1.431 ± 0.011 7.4 ± 0.6 171.7 ± 43.4

3.3. Electrochemical Impedance Spectroscopy


Nyquist plots representing the data obtained with EIS for ground and polished samples of AZ31
and AZ61 alloys were analyzed applying equivalent circuits shown in Figure 4. These equivalent
circuits consist of resistance of the corrosion environment (solution) Rs , the resistance of layer of
corrosion products R1 and resistance of the magnesium alloy base material R2 . In some cases,
the inductance L was present in the equivalent circuit, according to the character of the obtained
Nyquist plot. Constant phase element (CPE) represents capacity formed between the corrosion
environment and the corrosion products created on materials surface or the corrosion products and
the magnesium alloy, respectively. Polarization resistance Rp values were calculated according to the
equations presented under the equivalent circuits in Figure 4.

Figure 4. Equivalent circuits used to the evaluation of the obtained Nyquist plots: (a) with an inductive
loop in the equivalent circuit, (b) serial connection in the equivalent circuit and (c) parallel connection
in the equivalent circuit (CPE: Constant phase element).

Figure 5 shows Nyquist plots characterizing electrochemical corrosion behavior of ground and
polished samples of AZ31 and AZ61 alloys. Immersion times of the alloys in the corrosion environment
were from 5 min to 168 h. In all the plots at different immersion times, semicircles at high and low
frequencies are present. Obtained impedance data were analyzed using EC-Lab software and best-fitted
using the appropriate equivalent circuit model (Figure 4). The determined resulting Rp values are
given in Table 5.

39
Metals 2017, 7, 465

Figure 5. Nyquist plots of ground (#1200) and polished (0.25 μm) samples of AZ31 and AZ61
magnesium alloys obtained in enriched HBSS+ (with Mg2+ and Ca2+ ions): (a) AZ31—0.25 μm,
(b) AZ31—#1200, (c) AZ61—0.25 μm and (d) AZ61—#1200.

Table 5. Polarization resistance values obtained from EIS (electrochemical impedance spectroscopy)
measurements performed in enriched HBSS+ (with Mg2+ and Ca2+ ions).

Rp (Ω·cm2 )
Samples
5 min 1h 2h 4h 8h 12 h
AZ31—#1200 5209 ± 531 3626 ± 773 3495 ± 40 3489 ± 101 3924 ± 36 5331 ± 217
AZ61—#1200 6256 ± 19 15,458 ± 31 16,501 ± 58 18,402 ± 68 15,509 ± 333 8800 ± 564
AZ31—0.25 μm 4805 ± 142 6604 ± 695 3544 ± 53 3711 ± 115 3962 ± 445 5102 ± 964
AZ61—0.25 μm 5916 ± 37 14,404 ± 207 12,277 ± 232 14,203 ± 33 9193 ± 280 15,471 ± 575
Samples 24 h 48 h 72 h 96 h 168 h -
AZ31—#1200 5201 ± 516 4163 ± 356 2667 ± 776 4949 ± 624 7481 ± 115 -
AZ61—#1200 27,005 ± 396 24,392 ± 768 22,903 ± 769 22,326 ± 850 15,691 ± 2500 -
AZ31—0.25 μm 4139 ± 478 4405 ± 706 3455 ± 171 3668 ± 92 4853 ± 532 -
AZ61—0.25 μm 9097 ± 501 9786 ± 1020 14,501 ± 1236 12,693 ± 126 10,759 ± 318 -

3.4. Characterization of Corrosion Products


Figure 6 shows the SEM images of the surface of the AZ31 and AZ61 alloys after EIS measurements
(168 h of immersion in enriched HBSS+ with Mg2+ and Ca2+ ions).

Figure 6. Cont.

40
Metals 2017, 7, 465

Figure 6. Morphology of the corrosion products present on the surface of tested samples after EIS
measurement (168 h of immersion in enriched HBSS+ with Mg2+ and Ca2+ ions): (a) AZ31—0.25 μm,
(b) AZ31—#1200, (c) AZ61—0.25 μm and (d) AZ61—#1200.

Chemical composition (Table 6) of corrosion products present on the surfaces of the tested samples
shows the presence of higher amount of oxygen and phosphorus indicating the presence of phosphates,
oxides and hydroxides of magnesium (or calcium respectively).

Table 6. Chemical composition of corrosion product on the surface of AZ31 and AZ61 magnesium
alloys after EIS measurements in enriched HBSS+ (with Mg2+ and Ca2+ ions).

Chemical Composition (wt %)


Samples
Mg Al Cl P O C Na Ca S
AZ31—#1200 7.3 0.3 0.3 15.3 42.8 5.5 0.4 28.0 0.1
AZ61—#1200 14.4 6.2 0.5 10.6 44.2 13.6 0.4 9.8 0.3
AZ31—0.25 μm 5.1 0.1 0.2 15.6 42.4 5.0 0.4 31.1 0.1
AZ61—0.25 μm 14.4 4.6 0.6 11.8 43.5 11.3 0.5 13.1 0.2

Comparing the surface morphology of AZ31 magnesium alloy with polished and ground surface
more pronounced cracking of the layer of corrosion products created on the sample surface can be seen
for the ground sample (Figure 6b). The surface of the samples was covered by corrosion products based
on MgO, Mg(OH)2 [43,44] and due to the used corrosion environment also phosphate-based corrosion
products [45]. While the amount of MgO and Mg(OH)2 products is comparable for both the sample
surfaces, the phosphate-based layer of corrosion products seems to be thicker for the ground sample
(cracked layer presented on the sample surface, below the clusters of MgO and Mg(OH)2 products).
Significantly smaller amount of MgO and Mg(OH)2 products were observed on the surface of
AZ61 magnesium alloy when compared to AZ31 magnesium alloy. Similarly, as in the case of AZ31
magnesium alloy also in the case of AZ61 magnesium alloy a layer of phosphate-based corrosion
products was observed on samples surface. Thicker layer (larger features and cracks) seems to appear
on the surface of ground AZ61 magnesium alloy comparing of the polished AZ61 magnesium alloy
(Figure 6c,d).
Figure 7 shows the SEM images of the surface of the AZ31 and AZ61 alloys after EIS measurements
(168 h of immersion in HBSS without Mg2+ and Ca2+ ions).

41
Metals 2017, 7, 465

Figure 7. Morphology of the surface after EIS (electrochemical impedance spectroscopy) measurements
(168 h of immersion in HBSS without Mg2+ and Ca2+ ions): (a) AZ31—0.25 μm, (b) AZ31—#1200,
(c) AZ61—0.25 μm and (d) AZ61—#1200.

Chemical composition (Table 7) of corrosion products present on the surfaces of the tested samples
from Figure 7 shows the presence of oxygen and phosphorus indicating the presence of phosphates,
oxides and hydroxides of magnesium.

Table 7. Chemical composition of corrosion product on the surface of AZ31 and AZ61 magnesium
alloys after EIS measurements in HBSS (without Mg2+ and Ca2+ ions).

Chemical Composition (wt %)


Samples
Mg Al Cl P O C Na
AZ31—#1200 31.4 4.1 0.8 7.7 43.2 12.1 0.7
AZ61—#1200 36.3 6.1 0.6 6.5 36.3 13.3 0.9
AZ31—0.25 μm 35.8 4.7 0.5 7.1 37.9 13.1 0.9
AZ61—0.25 μm 26.6 7.9 0.5 7.1 45.2 11.8 0.9

The layer of phosphate-based corrosion products was observed on all samples’ surfaces.
Thicker layer (larger features and cracks) seems to appear on the surface of polished AZ61 magnesium
alloy comparing of the ground AZ61 magnesium alloy (Figure 7c,d). In the phosphate-based layer on
the AZ31 magnesium alloy surface were regions without any visible corrosion products (Figure 7b).
On the polished surface of the AZ31 magnesium alloy (Figure 7a) were moreover present small amount
of MgO and Mg(OH)2 products.

42
Metals 2017, 7, 465

4. Discussion
The effect of the surface treatment (ground vs. polished surface) was observed during potentiodynamic
measurements on both types of tested magnesium alloys. In both the cases, more positive value
of corrosion potential, Ecorr (Table 4), was determined for the polished samples (0.25 μm) when
compared to the ground samples (#1200). The microstructure of the AZ61 magnesium alloy contains
a higher number of intermetallic phases (Mg17 Al12 and AlMn, Figure 2) than the microstructure
of the AZ31 magnesium alloy containing only AlMn particles (Figure 1). All of these intermetallic
phase particles have more positive potential [46–48] than the substitution solid solution (α-Mg) which
caused the formation of microcells which usually result in the acceleration of the corrosion process.
The observations are in agreement with [49] where a positive effect of decreasing surface roughness on
AZ91 alloy electrochemical corrosion properties was presented. However, authors in [49] performed
the EIS measurement only 2 h of exposure of the material in 0.5 wt % NaCl solution and cannot detect
the influence of evolution of the layer of corrosion products and its adhesion to the material surface.
The difference in the Ecorr values determined for ground and polished surface was more significant
in the case of AZ61 magnesium alloy (Figure 8). The grinding process results in the higher roughness
of the treated surface comparing to the polished surface resulting in a larger real surface area exposed
to the corrosion environment [49]. Ecorr reaches more negative value comparing to the polished surface
and therefore the ground surface can be considered as less stable from a thermodynamic point of view
than the polished surface. While only a small amount of intermetallic phase particles was present in
the case of the AZ31 magnesium alloy, the effect of the chemical heterogeneity and surface roughness
was smaller compared to the AZ61 magnesium alloy samples. Corrosion potential Ecorr of the AZ31
and AZ61 alloys determined from measurements in HBSS without Mg2+ and Ca2+ ions reported by
authors in [32] have more negative values than data obtained in enriched HBSS+ with Mg2+ and
Ca2+ ions (Figure 8). In the [32] was not observed pitting corrosion attack of the samples (no pitting
potential Epit was observed on the curves) which indicate higher reactivity of tested magnesium alloys
in enriched HBSS+ probably caused by the content of Mg2+ , Ca2+ and sulphate ions.

(a) (b)

Figure 8. Comparison of (a) Ecorr and (b) icorr of ground (#1200) and polished (0.25 μm) surface AZ31
and AZ61 magnesium alloys in enriched HBSS+ (with Mg2+ and Ca2+ ions) and in HBSS (without
Mg2+ and Ca2+ ions [32]).

While in the case of AZ31 magnesium alloy a positive effect of polishing on the material response
of the corrosion environment from the kinetic point of view can be considered, opposite behavior was
observed in the case of AZ61 magnesium alloy. Polishing of the surface of AZ31 alloy resulted in 50%
decrease in the corrosion current density (icorr ). However, the same treatment resulted in 50% increase
of the icorr in the case of the AZ61 alloy. However, this behavior can be also explained by the presence
on a large amount of intermetallic phase particles in the microstructure of AZ61 alloy and higher real
surface area after grinding process.

43
Metals 2017, 7, 465

Corrosion current density (icorr ) and the resulting corrosion rate (vcorr ) were higher in the corrosion
environment of HBSS without Mg2+ and Ca2+ ions reported in [32] (Figure 8). Comparing to the
presented data the differences in the values were mostly about ten times. This effect could be caused
by the presence or absence of especially Mg2+ ions in the solution. In the case of HBSS without Mg2+
and Ca2+ ions could play a role the concentration gradient [50–52]. Magnesium in the alloys reacts
with the corrosion environment to produce Mg2+ ions. In the beginning of the immersion of the
samples no Mg2+ ions are present in the solution. Thanks to the concentration gradient the Mg2+
ions migrate from the place with high concentration place of the ions (the surface of the samples) to
the place with low concentration of the ions (HBSS without Mg2+ and Ca2+ ions). This migration of
the Mg2+ ions supports reactions of the corrosion environment with the surface of AZ31 and AZ61
alloys. On the other hand, the concentration gradient of the enriched HBSS+ with Mg2+ and Ca2+
ions should be lower compared to the solution without the ions because of the primary presence of
the Mg2+ ions in the corrosion environment. The migration of the Mg2+ ions from the surface of the
alloys to the corrosion environment is not so fast (like in HBSS without Mg2+ and Ca2+ ions) which
reduces a number of the amount of reactions of the corrosion environment with the surface of the
magnesium alloys.
This theory is supported by the results presented in [33,34,36]. The authors describe the effect
of the chlorides and sulphate ions concentration on the corrosion resistance of magnesium alloys.
With the increasing content of these ions, the icorr values increased and vcorr , respectively. However,
when comparing icorr values (Figure 8) in HBSS and enriched HBSS+, this phenomenon did not
occur. Enriched HBSS+ contains more chloride and sulphate ions (Table 1) but icorr values are lower.
Corrosion behavior of magnesium alloys in corrosion environment with sulphate ions is usually
measured in sodium sulphate solution [36] where the absence of Mg2+ ions could cause higher
reactivity of the alloys.
Potentiodynamic measurements are only short-time measurements. All the measurements take
approximately 10 min (5 min of stabilization time and 5 min of measurement itself). Therefore,
the potentiodynamic measurements are affected by conditions at the beginning of the measurement
like the concentration of specific ions in the solution, concentration gradient, etc. Moreover, if the
corrosion potential (Epit ) appear, no Tafel region could be observed in the anodic branch of the
polarization curves [38].
From the long-time point of view the corrosion behavior of AZ31 and AZ61 alloys can be
characterized by EIS. Electrochemical corrosion behavior of the magnesium alloys in enriched HBSS+
with Mg2+ and Ca2+ ions represented by the evolution of polarization resistance is shown in Figure 9a.
While almost no influence of the surface finish was observed in the case of AZ31 magnesium alloy,
higher values of polarization resistance were observed for ground AZ61 magnesium alloy comparing
to the polished material state.

Figure 9. Polarization resistance Rp of ground (#1200) and polished (0.25 μm) surface of AZ31 and
AZ61 magnesium alloys: (a) in enriched HBSS+ (with Mg2+ and Ca2+ ions) and (b) in HBSS (without
Mg2+ and Ca2+ ions [32]).

44
Metals 2017, 7, 465

In the case of AZ31 magnesium alloy the values of polarization resistance oscillated around
4000 Ω·cm2 for all the times of measurement for both the material states. The similar resistivity of both
the material states can be connected with similar corrosion products (amount and character) observed
on material surfaces (Figure 6a,b). On the surfaces of both the analyzed samples was observed the
presence of a layer of magnesium phosphate (Mg3 (PO4 )2 ) and hydroxiapatite (Ca10 (PO4 )6 (OH)2 )
primarily created on the sample surface and covered by large number of clusters of MgO and Mg(OH)2
products [45,53]. The assumption that the layer of the magnesium phosphate (Mg3 (PO4 )2 ) and
hydroxiapatite (Ca10 (PO4 )6 (OH)2 ) is thicker (Figure 6) in the case of ground sample correlate with
slightly higher values of Rp determined for the material state (Table 5). However, material surface
finish did not show any significant influence on AZ31 magnesium alloy corrosion resistivity in enriched
HBSS+ with Mg2+ and Ca2+ ions from the long-time point of view.
In the case of ground AZ61 magnesium alloy an increase of Rp up to 24 h of exposure followed
by its decrease back to the value characteristic for the beginning of the experiment was observed
(Table 5 and Figure 8a). In the case of polished AZ61 magnesium alloy a decrease of Rp up to 24 h
of exposure, it follows an increase up to 72 h of exposure followed by an additional decrease to the
value lower than the resistance of material on the beginning of the exposure was observed (Table 5 and
Figure 8). Higher values of Rp comparing to the AZ31 magnesium alloy can be explained by material
higher chemical heterogeneity and thus faster evolution of corrosion products on the material surface.
On the other hand, the differences between the polarization resistance of the ground and polished
AZ61 magnesium alloy can be corresponding to the higher material surface roughness. Higher real
surface area of the rough ground samples can be connected by faster growth of the layer of magnesium
phosphate (Mg3 (PO4 )2 ) and hydroxiapatite (Ca10 (PO4 )6 (OH)2 ) on sample surface comparing to the
polished sample surface. This was also observed in Figure 6c,d, where the thicker layer of corrosion
products was assumed for the ground sample when compared to the polished sample. The not stable
evolution of the Rp in time can correlate with the thicker layer of corrosion products and its cracking,
corrosion or remove from the samples surface, which was observed only in a minor form in the case of
AZ31 magnesium alloy with the stable evolution of Rp in time. The presence of inductive loop L [54,55]
in the equivalent circuit implies the occurrence of pitting corrosion on the magnesium alloy’s surface.
The values of polarization resistance Rp are quite high at the beginning of exposition of the
samples in the corrosion environment. The values of Rp support the theory of Mg2+ ions concentration
gradient as well as in the case of explanation for icorr change (vcorr respectively). Magnesium reacts
more slowly when Mg2+ ions are present in the corrosion environments as itself (enriched HBSS+ with
Mg2+ and Ca2+ ions). Helmholtz double layer created on the interface magnesium alloy/corrosion
environment is more stable as in the case of HBSS without Mg2+ ions and the resistance of the material
is high.
The opposite phenomenon can be observed in the case of HBSS without Mg2+ and Ca2+ ions.
At the beginning of the exposition of the samples to this corrosion environment the Rp values are lower
(Figure 9b) than in the case of the enriched HBSS+ with Mg2+ and Ca2+ ions and the value increase
with increasing time of exposure to the corrosion environment.
Up to 24 h of exposure comparable values of Rp were determined for both the alloys and for both
the material states in HBSS without Mg2+ and Ca2+ ions (Figure 9b). With increasing exposure time
the values of Rp slowly decreased for AZ31 magnesium alloy, while slightly higher values of Rp were
determined for the ground samples. This observation corresponds to the smaller number of cracks in
the layer of corrosion products present on the surface of ground samples comparing to the polished
samples (Figure 7a,b). An increase of Rp above 24 h of exposure to the corrosion environment was
observed for AZ61 magnesium alloy, while the highest values were determined for 168 h of exposure
to the HBSS without Mg2+ and Ca2+ ions (Figure 9b). Same as in the case of the AZ31 alloy, also in the
case of AZ61 alloy slightly higher values of Rp were determined for the ground samples. Figure 7c,d
showing the surface of the samples after the EIS measurements can exhibit the explanation for this
material behavior. While the smaller number of finer cracks were observed on the surface of the layer

45
Metals 2017, 7, 465

of corrosion products created on the ground sample compared to the deep and large cracks observed
on the surface of the polished sample.
In the both corrosion environments, the influence of corrosion products must be taken into
account when predicting material corrosion behavior, because magnesium and its alloys are very
reactive by themselves. The corrosion layer behaves as a barrier against the further access of corrosion
environment to the surface of the magnesium alloys. The layers of the corrosion products are composed
mainly of the oxides, hydroxides, phosphates and chlorides [45]. However, the layer is porous and
uncompacted (cracks are present in Figures 6 and 7) which leads to other reactions in the corrosion
environments of the samples. Even though, different corrosion products are created on the surface of
AZ31 and AZ61 alloys due to the chemical composition of the used corrosion solution, more stable
material behavior was observed for the AZ31 magnesium alloy samples in both of the solutions.
AZ61 magnesium alloy exhibited more stable corrosion response in the HBSS without Mg2+ and Ca2+
ions [32] comparing to its behavior in enriched HBSS+ with Mg2+ and Ca2+ ions.

5. Conclusions
This work was focused on the evaluation of corrosion behavior of AZ31 and AZ61 magnesium
alloys in enriched HBSS+ (with Mg2+ and Ca2+ ions) by electrochemical methods and the results were
compared with results presented in [32]. The results of the work can be summarized as follows:

(1) Electrochemical corrosion properties of AZ31 and AZ61 alloys are dependent on corrosion
environment composition.
(2) AZ31 and AZ61 alloys had more positive corrosion potential Ecorr and significantly lower values
of corrosion current density icorr in the enriched HBSS+ (with Mg2+ and Ca2+ ions) comparing to
the standard HBSS (without Mg2+ and Ca2+ ions). It was also observed that the pitting corrosion
attack with pitting potential Epit in the enriched HBSS+ solution (with Mg2+ and Ca2+ ions)
which resulted in less uniform corrosion attack (less predictable material behavior) of the material
compared to material behavior in HBSS.
(3) No significant influence of the surface treatment (ground vs. polished) was observed during EIS
measurements of AZ31 magnesium alloy performed in both types of the HBSSs.
(4) Polarization resistance Rp of AZ61 magnesium alloy in enriched HBSS+ (with Mg2+ and Ca2+
ions) was affected by surface treatment. Rp values of ground samples were higher than Rp of
polished samples, after 24 h of exposure.
(5) AZ61 magnesium alloy reached higher Rp values than AZ31 magnesium alloy in both the used
HBSSs which indicated better corrosion resistivity.
(6) While only phosphate-based corrosion products layer was characteristic for magnesium alloys
EIS tested in HBSS, the combination of phosphate-based corrosion products layer and clusters
of MgO and Mg(OH)2 products were presented on the surface of specimens tested in enriched
HBSS+ (with Mg2+ and Ca2+ ions).

Acknowledgments: This work was supported by the project “Materials Research Centre at FCH BUT—
Sustainability and Development”, REG LO1211, with financial support from National Programme for
Sustainability I (Ministry of Education, Youth and Sports), Czech Republic; NETME center plus, projects of
Ministry of Education, Youth and Sports of the Czech Republic under the “National sustainability program”;
and by the Slovak Research and Development Agency for support in experimental by the projects
No. APVV-14-0284, No. APVV-14-0772 and No. APVV-14-0096.
Author Contributions: Jakub Tkacz, Pavel Doležal and Jaromír Wasserbauer conceived and designed the
experiments; Karolína Slouková and Jozef Minda performed electrochemical measurements; Jakub Tkacz and
Juliána Drábiková performed SEM analysis; Pavel Doležal provided experimental materials; Jakub Tkacz and
Stanislava Fintová wrote the paper.
Conflicts of Interest: The authors declare no conflict of interest.

46
Metals 2017, 7, 465

References
1. Jung, O.; Smeets, R.; Porchetta, D.; Kopp, A.; Ptock, C.; Müller, U.; Heiland, M.; Schwade, M.; Behr, B.;
Kröger, N.; et al. Optimized in vitro procedure for assessing the cytocompatibility of magnesium-based
biomaterials. Acta Biomater. 2015, 23, 354–363. [CrossRef] [PubMed]
2. Farraro, K.F.; Kim, K.E.; Woo, S.L.; Flowers, J.R.; McCullough, M.B. Revolutionizing orthopedic biomaterials:
The potential of biodegradable and bioresorbable magnesium-based materials for functional tissue
engineering. J. Biomech. 2014, 47, 1979–1986. [CrossRef] [PubMed]
3. Mhaede, M.; Pastorek, F.; Hadzima, B. Influence of shot peening on corrosion properties of biocompatible
magnesium alloy AZ31 coated by dicalcium phosphate dihydrate (DCPD). Mater. Sci. Eng. C 2014, 39,
330–335. [CrossRef] [PubMed]
4. Gu, X.N.; Zheng, Y.F. A review on magnesium alloys as biodegradable materials. Front. Mater. Sci. China
2010, 4, 111–115. [CrossRef]
5. Witte, F.; Kaese, V.; Haferkamp, H.; Switzer, E.; Meyer-Lindenberg, A.; Wirth, C.J.; Windhagen, H. In vivo
corrosion of four magnesium alloys and the associated bone response. Biomaterials 2005, 26, 3557–3563.
[CrossRef] [PubMed]
6. Li, W.; Guan, S.; Chen, J.; Hu, J.; Chen, S.; Wang, L.; Zhu, S. Preparation and in vitro degradation of the
composite coating with high adhesion strength on biodegradable Mg-Zn-Ca alloy. Mater. Charact. 2011, 62,
1158–1165. [CrossRef]
7. Homayun, B.; Afshar, A. Microstructure, mechanical properties, corrosion behavior and cytotoxicity of
Mg-Zn-Al-Ca alloys as biodegradable materials. J. Alloys Compd. 2014, 607, 1–10. [CrossRef]
8. Bakhsheshi-Rad, H.R.; Idris, M.H.; Abdul-Kadir, M.R.; Ourdjini, A.; Medraj, M.; Daroonparvar, M.;
Hamzah, E. Mechanical and bio-corrosion properties of quaternary Mg-Ca-Mn-Zn alloys compared with
binary Mg-Ca alloys. Mater. Des. 2014, 53, 283–292. [CrossRef]
9. Staiger, M.P.; Pietak, A.M.; Huadmai, J.; Dias, G. Magnesium and its alloys as orthopedic biomaterials:
A review. Biomaterials 2006, 27, 1728–1734. [CrossRef] [PubMed]
10. Song, G. Control of biodegradation of biocompatable magnesium alloys. Corros. Sci. 2007, 49, 1696–1701.
[CrossRef]
11. Salleh, E.M.; Zuhailawati, H.; Ramakrishnan, S.; Gepreel, M.A.-H. A statistical prediction of density
and hardness of biodegradable mechanically alloyed Mg-Zn alloy using fractional factorial design.
J. Alloys Compd. 2015, 644, 476–484. [CrossRef]
12. Plum, L.M.; Rink, L.; Haase, H. The essential toxin: Impact of zinc on human health. Int. J. Environ. Res.
Public Health 2010, 7, 1342–1365. [CrossRef] [PubMed]
13. Atkins, P. Physical Chemistry, 8th ed.; Freeman and Company: New York, NY, USA, 2006; pp. 1–1053.
14. Gray, J.E.; Luan, B. Protective coatings on magnesium and its alloys—A critical review. J. Alloys Compd. 2002,
336, 88–113. [CrossRef]
15. Song, G.L.; Atrens, A. Corrosion mechanism of magnesium alloys. Adv. Eng. Mater. 1999, 1, 11–13. [CrossRef]
16. Ferero López, A.D.; Lehr, I.L.; Saidman, S.B. Anodisation of AZ91D magnesium alloy in molybdate solution
for corrosion protection. J. Alloys Compd. 2017, 702, 338–345. [CrossRef]
17. Kannan, M.B.; Raman, R.K.S. In vitro degradation and mechanical integrity of calcium-containing
magnesium alloys in modified-simulated body fluid. Biomaterials 2008, 29, 2306–2314. [CrossRef] [PubMed]
18. Xin, Y.; Hu, T.; Chu, P.K. In vitro studies of biomedical magnesium alloys in a simulated physiological
environment: A review. Acta Biomater. 2011, 7, 1452–1459. [CrossRef] [PubMed]
19. Proudfoot, A.T. Aluminium and zinc phosphide poisoning. Clin. Toxicol. 2009, 47, 89–100. [CrossRef]
[PubMed]
20. Seitz, J.M.; Eifler, R.; Bach, F.W.; Maier, H.J. Magnesium degradation products: Effects on tissue and human
metabolism. J. Biomed. Mater. Res. A 2014, 102, 3744–3753. [CrossRef] [PubMed]
21. Adeknmbi, I.; Mosher, C.Z.; Lu, H.H.; Reihle, M.; Kubba, H.; Tanner, K.E. Mechanical behaviour of
biodegradable AZ31 magnesium alloy after long term in vitro degradation. Mater. Sci. Eng. C 2017,
77, 1135–1144. [CrossRef] [PubMed]
22. Li, Q.; Jiang, G.; Wang, C.; Dong, J.; He, G. Mechanical degradation of porous titanium with entangled
structure filled with biodegradable magnesium in Hanks’ solution. Mater. Sci. Eng. C 2015, 57, 349–354.
[CrossRef] [PubMed]

47
Metals 2017, 7, 465

23. Brar, H.S.; Wong, J.; Manuel, M.V. Investigation of the mechanical and degradation properties of Mg-Sr and
Mg-Zn-Sr alloys for use as potential biodegradable implant materials. J. Mech. Behav. Biomed. Mater. 2012, 7,
87–95. [CrossRef] [PubMed]
24. Berglund, I.S.; Brar, H.S.; Dolgova, N.; Acharya, A.P.; Keselowsky, B.G.; Sarntinoranont, M.; Manuel, M.V.
Synthesis and characterization of Mg-Ca-Sr alloys for biodegradable orthopedic implant applications.
J. Biomed. Mater. Res. Part B Appl. Biomater. 2012, 100, 1524–1534. [CrossRef] [PubMed]
25. Zhang, F.; Ma, A.; Song, D.; Jiang, J.; Lu, F.; Zhang, L.; Yang, D.; Chen, J. Improving in vitro biocorrosion
resistance of Mg-Zn-Mn-Ca alloy in Hank’s solution through addition of cerium. J. Rare Earths 2015, 33,
93–101. [CrossRef]
26. Johnston, S.; Shi, Z.; Atrens, A. The influence of pH on the corrosion rate of high-purity Mg, AZ91 and ZE41
in bicarbonate buffered Hanks’ solution. Corros. Sci. 2015, 101, 182–192. [CrossRef]
27. Zainal Abidin, N.I.; Rolfe, B.; Owen, H.; Malisano, J.; Martin, D.; Hofstetter, J.; Uggowitzer, P.J.; Atrens, A.
The in vivo and in vitro corrosion of high-purity magnesium and magnesium alloys WZ21 and AZ91.
Corros. Sci. 2013, 75, 354–366. [CrossRef]
28. Han, G.; Lee, J.-Y.; Kim, Y.-C.; Park, J.H.; Kim, D.-I.; Han, H.-S.; Yang, S.-J.; Seok, H.-K.
Preferred crystallographic pitting corrosion of pure magnesium in Hanks’ solution. Corros. Sci. 2012,
63, 316–322. [CrossRef]
29. Zeng, R.-C.; Sun, L.; Zheng, Y.-F.; Cui, H.-Z.; Han, E.-H. Corrosion and characterisation of dual phase
Mg-Li-Ca alloy in Hank’s solution: The influence of microstructural features. Corros. Sci. 2014, 79, 69–82.
[CrossRef]
30. Ng, W.F.; Chiu, K.Y.; Cheng, F.T. Effect of pH on the in vitro corrosion rate of magnesium degradable implant
material. Mater. Sci. Eng. C 2010, 30, 898–903. [CrossRef]
31. Bukovinová, L.; Hadzima, B. Electrochemical characteristics of magnesium alloy AZ31 in Hank’s solution.
Corros. Eng. Sci. Technol. 2012, 47, 352–357. [CrossRef]
32. Tkacz, J.; Slouková, K.; Minda, J.; Drábiková, J.; Fintová, S.; Doležal, P.; Wasserbauer, J. Corrosion behavior of
wrought magnesium alloys AZ31 and AZ61 in Hank’s solution. Koroze a Ochrana Materiálu 2016, 60, 101–106.
[CrossRef]
33. Ambat, R.; Aung, N.N.; Zhou, W. Studies on the influence of chloride ion and pH on the corrosion and
electrochemical behaviour of AZ91D magnesium alloy. J. Appl. Electrochem. 2000, 30, 865–874. [CrossRef]
34. Altun, H.; Sen, S. Studies on the influence of chloride ion concentration and pH on the corrosion and
electrochemical behaviour of AZ63 magnesium alloy. Mater. Des. 2004, 25, 637–643. [CrossRef]
35. Merino, M.C.; Pardo, A.; Arrabal, R.; Merino, S.; Casajus, P.; Mohedano, M. Influence of chloride ion
concentration and temperature on the corrosion of Mg-Al alloys in salt fog. Corros. Sci. 2010, 52, 1696–1704.
[CrossRef]
36. Shetty, S.; Nayak, J.; Shetty, N. Influence of sulfate ion concentration and pH on the corrosion of Mg-Al-Zn-Mn
(GA9) magnesium alloy. J. Magnes. Alloys 2015, 3, 258–270. [CrossRef]
37. Vander Voort, G.F. Metallography, Principles and Practice, 1st ed.; ASM International: Materials Park, OH, USA,
1999; pp. 1–752.
38. Tkacz, J.; Minda, J.; Fintová, S.; Wasserbauer, J. Comparison of electrochemical methods for the evaluation of
cast AZ91 magnesium alloy. Materials 2016, 9, 925. [CrossRef] [PubMed]
39. Liu, F.; Song, Y.; Shan, D.; Han, E. Corrosion behavior of AZ31 magnesium alloy in simulated acid rain
solution. Trans. Nonferr. Met. Soc. China 2010, 20, 638–642. [CrossRef]
40. Deng, J.; Huang, G.; Zhao, Y.; Wang, B. Electrochemical performance of AZ31 magnesium alloy under
different processing conditions. Rare Met. Mater. Eng. 2014, 43, 316–321.
41. Standard Specification for Magnesium-Alloy Sheet and Plate; ASTM B90/B90M-15; ASTM International:
West Conshohocken, PA, USA, 2015. Available online: www.astm.org (accessed on 20 October 2017).
42. Standard Specification for Magnesium-Alloy Extruded Bars, Rods, Profiles, Tubes, and Wire; ASTM B107/B107M-13;
ASTM International: West Conshohocken, PA, USA, 2013. Available online: www.astm.org (accessed on
20 October 2017).
43. Harandi, S.E.; Banerjee, P.C.; Easton, C.D.; Raman, R.K.S. Influence of bovine serum albumin in Hanks’
solution on the corrosion and stress corrosion cracking of a magnesium alloy. Mater. Sci. Eng. C 2017, 80,
335–345. [CrossRef] [PubMed]

48
Metals 2017, 7, 465

44. Zhu, Y.; Wu, G.; Zhang, Y.H.; Zhao, Q. Growth and characterization of Mg(OH)2 film on magnesium alloy
AZ31. Appl. Surf. Sci. 2011, 257, 6129–6137. [CrossRef]
45. Song, Y.; Shan, D.; Chen, R.; Zhang, F.; Han, E.-H. Biodegradable behaviors of AZ31 magnesium alloy in
simulated body fluid. Mater. Sci. Eng. C 2009, 29, 1039–1045. [CrossRef]
46. Mathieu, S.; Rapin, C.; Steinmetz, J.; Steinmetz, P. A corrosion study of the main constituent phases of AZ91
magnesium alloys. Corros. Sci. 2003, 45, 2741–2755. [CrossRef]
47. Pardo, A.; Merino, M.C.; Coy, A.E.; Arrabal, R.; Viejo, F.; Matykina, E. Corrosion behaviour of
magnesium/aluminium alloys in 3.5 wt % NaCl. Corros. Sci. 2008, 50, 823–834. [CrossRef]
48. Feliu, S.; Llorente, I. Corrosion product layers on magnesium alloys AZ31 and AZ61: Surface chemistry and
protective ability. Appl. Surf. Sci. 2015, 347, 736–746. [CrossRef]
49. Walter, R.; Kannan, M.B. Influence of surface roughness on the corrosion behaviour of magnesium alloy.
Mater. Des. 2011, 32, 2350–2354. [CrossRef]
50. Esmaily, M.; Svensson, J.E.; Fajardo, S.; Birbilis, N.; Frankel, G.S.; Virtanen, S.; Arrable, R.; Thomas, S.;
Johansson, L.G. Fundamentals and advances in magnesium alloy corrosion. Prog. Mater. Sci. 2017, 89, 92–193.
[CrossRef]
51. Zhou, M.; Yan, L.; Ling, H.; Diao, Y.; Pang, X.; Wang, Y. Design and fabrication of enhanced corrosion
resistance Zn-Al layered double hydroxides films based anion-exchange mechanism on magnesium alloys.
Appl. Surf. Sci. 2017, 404, 246–253. [CrossRef]
52. Tian, J.; Huang, H.L.; Pan, Z.Q.; Zhou, H. Effect of flow velocity on corrosion behavior of AZ91D magnesium
alloy at elbow of loop system. Trans. Nonferr. Met. Soc. China 2016, 26, 2857–2867. [CrossRef]
53. Zhu, B.; Wang, S.; Wang, L.; Yang, Y.; Liang, J.; Cao, B. Preparation of hydroxyapatite/tannic acid coating to
enhance the corrosion resistance and cytocompatibility of AZ31 magnesium alloys. Coatings 2017, 7, 105.
[CrossRef]
54. Jayaraj, J.; Raj, S.A.; Srinivasan, A.; Ananthakumar, S.; Pillay, U.T.S.; Dhaipule, N.G.K.; Mudali, U.K.
Composite magnesium phosphate coatings for improved corrosion resistance of magnesium AZ31 alloy.
Corros. Sci. 2016, 113, 104–115. [CrossRef]
55. Yagi, S.; Sengoku, A.; Kubota, K.; Matsubara, E. Surface modification of ACM522 magnesium alloy by
plasma electrolytic oxidation in phosphate electrolyte. Corros. Sci. 2012, 57, 74–80. [CrossRef]

© 2017 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

49
metals
Article
Characterization of Powder Metallurgy Processed
Pure Magnesium Materials for
Biomedical Applications
Matěj Březina 1, *, Jozef Minda 1 , Pavel Doležal 1,2 , Michaela Krystýnová 1 , Stanislava Fintová 1,3 ,
Josef Zapletal 2 , Jaromír Wasserbauer 1 and Petr Ptáček 1
1 Materials Research Centre, Faculty of Chemistry, Brno University of Technology, Purkyňova 464/118,
61200 Brno, Czech Republic; xcminda@fch.vut.cz (J.M.); dolezal@fme.vutbr.cz (P.D.);
xckrystynovam@fch.vut.cz (M.K.); fintova@ipm.cz (S.F.); wasserbauer@fch.vut.cz (J.W.);
ptacek@fch.vut.cz (P.P.)
2 Institute of Materials Science and Engineering, Faculty of Mechanical Engineering,
Brno University of Technology, Technická 2896/2, 61669 Brno, Czech Republic; zapletal@fme.vutbr.cz
3 Institute of Physics of Materials, Academy of Sciences of the Czech Republic, Žižkova 22,
61662 Brno, Czech Republic
* Correspondence: xcbrezinam@fch.vutbr.cz; Tel.: +420-54-114-9469

Received: 14 September 2017; Accepted: 24 October 2017; Published: 31 October 2017

Abstract: Magnesium with its mechanical properties and nontoxicity is predetermined as a


material for biomedical applications; however, its high reactivity is a limiting factor for its
usage. Powder metallurgy is one of the promising methods for the enhancement of material
mechanical properties and, due to the introduced plastic deformation, can also have a positive
influence on corrosion resistance. Pure magnesium samples were prepared via powder metallurgy.
Compacting pressures from 100 MPa to 500 MPa were used for samples’ preparation at room
temperature and elevated temperatures. The microstructure of the obtained compacts was analyzed
in terms of microscopy. The three-point bending test and microhardness testing were adopted
to define the compacts’ mechanical properties, discussing the results with respect to fractographic
analysis. Electrochemical corrosion properties analyzed with electrochemical impedance spectroscopy
carried out in HBSS (Hank’s Balanced Salt Solution) and enriched HBSS were correlated with the
metallographic analysis of the corrosion process. Cold compacted materials were very brittle with
low strength (up to 50 MPa) and microhardness (up to 50 HV (load: 0.025 kg)) and degraded
rapidly in both solutions. Hot pressed materials yielded much higher strength (up to 250 MPa)
and microhardness (up to 65 HV (load: 0.025 kg)), and the electrochemical characteristics were
significantly better when compared to the cold compacted samples. Temperatures of 300 ◦ C and
400 ◦ C and high compacting pressures from 300 MPa to 500 MPa had a positive influence on material
bonding, mechanical and electrochemical properties. A compacting temperature of 500 ◦ C had a
detrimental effect on material compaction when using pressure above 200 MPa.

Keywords: magnesium; powder metallurgy; cold pressing; hot pressing; EIS (Electrochemical
impedance spectroscopy); three-point bending test; corrosion

1. Introduction
Magnesium and its alloys are modern lightweight materials applicable in a wide range of
industrial fields from aerospace and automotive to biomedical applications. Its main advantages are a
good strength to weight ratio and biocompatibility in combination with biodegradability. However,
due to the high reactivity of pure Mg and the mechanical properties, not really sufficient for engineering
applications, mainly magnesium alloys are used [1–4].

Metals 2017, 7, 461; doi:10.3390/met7110461 50 www.mdpi.com/journal/metals


Metals 2017, 7, 461

Good mechanical properties of magnesium and its alloys can be furthermore significantly
upgraded by decreasing the grain size, nowadays performed mainly via severe plastic deformation
(SPD) techniques [3,5], powder metallurgy (PM) processing [1,2,6,7] or by a combination of both
methods. Decreasing the metallic grain size within the material volume, either by alloying elements,
SPD methods, such as extrusion, equal channel angular pressing (ECAP) and high-pressure torsion
(HPT) or PM, leads to increase of hardness, tensile and yield strength, but the plasticity of the material
was shown to be decreased [6–10].
PM processing of magnesium is influenced by its high affinity to oxygen, which results in the
formation of a thermodynamically-stable layer of corrosion products on the magnesium powder
particles’ surface. The created layer dramatically inhibits the diffusion processes required for material
densification during PM processing [4]. Because of the high affinity of magnesium to oxygen,
a protective atmosphere (usually argon or nitrogen) has to be used for handling of magnesium
powders and specimens, as well as for subsequent sintering [4].
Applying high pressures at elevated temperatures (hot pressing) to process magnesium results in
high plastic deformation of powder particles. Powder particles’ deformation leads to cracking of the
layer of corrosion products normally present on the powder particles’ surface [11–13]. The applied
plastic deformation results in an increase of the contact area of powder particles and, in combination
with the applied temperature, enhances the diffusion processes. Increasing compacting pressures
usually also lead to a decrease in the porosity of the processed bulk material [11–13]. The porosity of
the processed bulk material is usually considered as a disadvantage of the PM processing techniques.
However, the porous biocompatible material can be incorporated well into the tissue and can degrade
at a specific rate, which provides a tool for the tailoring properties of PM processed materials for
biomedical applications [14,15]. The functional porosity of a PM processed magnesium-based implant
would support the primary fixation and the degradation of the implant by enabling the ingrowth
of bone cells (osteointegration) into the degrading implant. Furthermore, the corrosion products of
magnesium, created during implant biodegradation, support osteoconductivity of the bone [4].
Changes in porosity also have a significant effect on corrosion resistance and corrosion attack
progression within the material volume. Highly porous materials corrode very rapidly, as a larger
area of the material surface is exposed to the corrosion environment. Corrosion resistance of either
pure magnesium or magnesium alloys is seldom suitable for technical applications or even biomedical
applications [3,9,16–20]. Magnesium corrosion resistance can be improved by alloying the metal with
aluminum, zinc or rare earth metal elements; however, for significantly better corrosion resistance,
another way of reducing the degradation rate must be considered. Conversion coatings are widely
studied as corrosion protection for magnesium and its alloys. Fluoride and calcium phosphate-based
conversion coatings have a great potential of reducing the corrosion rate of biomedical magnesium
implants [1–4,12–24].
Electrochemical characterization of PM magnesium materials is commonly carried out either in
NaCl solution or in Hank's Balanced Salt Solution (HBSS) or enriched HBSS [21,25–29]. Electrochemical
impedance spectroscopy (EIS) measurements provide complex information of the material degradation
characteristics during exposure in corrosion medium in time. From the chemical point of view, several
specific reactions can occur in HBSS and enriched HBSS medium due to the additional content of
several ions when compared to NaCl solution [30].
The properties of pure magnesium materials prepared via PM methods are seldom
studied [15,27,31–33] or the studies refer to wrought pure magnesium materials [5]. Most of the available
studies characterize the PM processed pure magnesium and magnesium-based materials in terms of
mechanical and corrosion properties; however, the analyses are focused on the influence of the material
porosity due to its possible use for biomedical applications in the form of the porous implants.
Porous magnesium samples with different porosity (due to the addition of an ammonium
bicarbonate powder) structures and mechanical properties were analyzed in [31]. Magnesium powder
samples with different porosities, due to the spacer addition, were prepared by uniaxial cold pressing

51
Metals 2017, 7, 461

(265 MPa) with additional heat treatment (for removal of hexane and sintering at 550 ◦ C for 6 h). In the
case of pure magnesium compacts, the porosity of 12 vol % corresponded to a flexural strength of
38 MPa. The increase of the compacts’ porosity was observed to be accompanied by the mechanical
properties’ decrease (the porosity of 38 vol % corresponded to the flexural strength of 4.4 MPa).
The negative influence of the porosity was observed also in the case of material immersion into 9 g/L
NaCl solution.
The same authors studied also the influence of the purity of the argon atmosphere on final PM
processed magnesium-based material properties in [32], varying the sintering times as 0 h, 3 h, 6 h, 12 h
and 24 h. While the samples prepared using technical Ar reached similar porosities in the sintering
time range, the porosity of the samples prepared using gettered Ar was observed to decrease with
increasing sintering time. Under a gettered Ar atmosphere, a prolonged sintering time enhanced
diffusion connections between magnesium particles and improved the mechanical properties of the
samples (flexural strength of 9 MPa for 3 h and 115.4 MPa for 24 h), whereas under a technical Ar
atmosphere (flexural strength of 5 MPa for 3 h and 3.4 MPa for 24 h), oxidation at the particle surfaces
caused deterioration in the mechanical properties of the samples.
Mechanical properties and corrosion resistance of cold pressed magnesium (310 MPa) and
extruded at 420 ◦ C employing an extrusion ratio of 16:1 were studied in [24]. Processed material
reached the ultimate tensile strength of 320 MPa, a yield stress of 280 MPa and 2% ductility. The reached
values were higher when compared to a coarse-grained cast AZ31 magnesium alloy, except material
ductility. PM processed pure magnesium mechanical properties were superior to those of cast
pure magnesium as provided by the authors. Silane film and anticorrosive paint were shown to
enhance the corrosion behavior of PM magnesium during the first hours of immersion, but their
protection effectiveness completely disappears after two days. For longer immersion times, the fluoride
conversion coating prepared in HF solution was shown to be an effective barrier to protect PM
magnesium from degradation.
Pre-rolling of magnesium powder prior to spark plasma sintering and hot extrusion was used for
material grain refinement in [6]. The grain size obtained for the material extruded from the powders
after 5 and 10 rolling pre-treatments decreased from 9.2 μm (only extruded Mg) to 2.9 μm and 2.1 μm,
respectively. The grain refinement resulted in an ultimate tensile strength increase from 242 MPa for
extruded pure magnesium to a value of 270.6 MPa for material extruded after sintering of 10 times
pre-rolled powder. Yield stress was improved from 170 MPa to 206.4 MPa by the pre-rolling of powder.
The aim of this study is to evaluate the microstructural, mechanical and electrochemical
properties of pure magnesium-based PM processed materials when varying the processing parameters.
Materials were prepared via hot pressing at temperatures of 300 ◦ C, 400 ◦ C and 500 ◦ C applying
pressures of 100 MPa, 200 MPa, 300 MPa, 400 MPa and 500 MPa. For comparison, cold compacted
materials prepared under the same compaction pressures were analyzed in the same manner.
The mechanical properties of compacts were analyzed in terms of the three-point bending test
containing fractographic analysis and with microhardness testing. Electrochemical corrosion
characteristics of the processed materials in HBSS and enriched HBSS were analyzed using
electrochemical impedance spectroscopy, and the data were extended by metallographic analysis
of the corrosion attack within the material volume.

2. Materials and Methods


Magnesium powder used in this study (Figure 1) was irregularly shaped with an average particle
size of approximately 30 μm. The purity of the base material was 99.8% as declared by supplier
Goodfellow (Huntingdon, UK); however, an oxide layer was found on the surface of powder particles
using the scanning electron microscope (SEM, ZEISS EVO LS 10, Carl Zeiss Ltd., Cambridge, UK)
using energy dispersive spectroscopy (EDS, Oxford Instruments plc, Abingdon, UK). This layer is to
be expected on the surface of magnesium, and it was probably presented on the particles from the
powder preparation.

52
Metals 2017, 7, 461

Figure 1. Magnesium powder SEM (scanning electron microscope): powder particles’ morphology
and size distribution.

For metallographic and electrochemical analysis, cylindrical compacts of 5 mm in height and


20 mm in diameter were prepared using a steel die. The preparation of the magnesium powder
into the steel die for compaction was carried out under a nitrogen atmosphere to avoid further
oxygen contamination. Magnesium powder (2.7 g) inserted into the steel die was compacted into
compacts applying different uniaxial pressures; 100 MPa, 200 MPa, 300 MPa, 400 MPa and 500 MPa.
Compaction was carried out using the Zwick Z250 Allround-Line machine (Zwick GmbH&Co. KG in
Germany, Ulm, Germany) at room temperature and at elevated temperatures of 300 ◦ C, 400 ◦ C and
500 ◦ C for one hour.
The microstructure of prepared samples was studied on cross-sectional cuts using SEM (ZEISS
EVO LS 10) and a light optical microscope (LM, Zeiss Axio Observer Z1m, Carl Zeiss AG,
Oberkochen, Germany). The internal structure of the compacts and the created grain size were
analyzed with the electron backscattered diffraction (EBSD) technique (SEM, Tescan LYRA 3 XMU
FEG/SEMxFIB, TESCAN Brno, s.r.o., Brno, Czech Republic). The step size of the method was 0.3 μm.
Samples’ cross-sections intended for metallographic analysis were mold into a polymeric resin at room
temperature, ground and polished according to the standard procedures of metallographic samples’
preparation. To reveal the microstructure, a 5% Nital etchant (5% nitric acid in ethanol) was used for
several seconds. Compacts’ porosity was estimated from calculated density (calculated from sample
dimensions and weight) considering a density of cast pure magnesium as the 0 porosity standard.
Microhardness testing was conducted on cross-sections of prepared compacts using the LECO
AMH LM 248 microhardness tester (LECO Corporation, Saint Joseph, MI, USA) according to the ISO
6507-1 standard. For all the measurement, 25 g of load and a 10-s dwell time were used. The three-point
bending test was carried out on samples with dimensions of 4 mm in height, 4 mm in width and
18 mm in length cut from the compacts, according to the ISO 7438 standard. The test machine used for
the three-point bending test was the Zwick Z020 (Zwick GmbH&Co. KG in Germany, Ulm, Germany);
the radius of the supports and the point was 2.5 mm; the distance between supports was 16 mm;
and the velocity of applied loading was 1 mm/min. The three-point bending test was performed on
one sample for each material preparation method.
Potentiostatic electrochemical impedance spectroscopy (EIS) was used as a method for the
characterization of the prepared magnesium samples’ electrochemical corrosion characteristics. The three
electrode cell system was used for the measurement with the Pt electrode as the counter electrode and
the calomel electrode as the reference electrode, and the sample (1 cm2 exposed area) served as the
working electrode. With the aim to remove a layer of corrosion products created on the compacts’
surface, each sample was ground using 4000 SiC paper and rinsed with isopropanol just before the
measurement. Measurements were carried out in HBSS and enriched HBSS (containing Mg2+ and Ca2+
ions). The composition of the used solutions is given in Table 1. The frequency used for the measurements
was in the range from 100 kHz to 10 MHz, with the signal amplitude of 10 mV. All the measurements
were carried out at laboratory temperature. EIS data were obtained after 5 min, 1 h, 2 h, 4 h, 8 h, 12

53
Metals 2017, 7, 461

h, 24 h, 48 h, 72 h and 96 h of immersion. Each measurement time was 25.8 min. Due to the open
porosity of the cold pressed samples, only cold pressed samples prepared under 500 MPa compacting
pressure were used for the electrochemical corrosion characterization. In the case of hot pressing, samples
prepared under 100 MPa and 500 MPa prepared at 400 ◦ C were used for EIS measurements with the aim
to reveal differences in material electrochemical corrosion behavior according to the processing conditions.
Each material type was characterized by measurements performed on three samples.

Table 1. Composition of the HBSS (Hank’s Balanced Salt Solution) and enriched HBSS (supplied by
GE (General Electric) Healthcare) used.

Concentration (mg·dm−3 )
Component
HBSS Enriched HBSS
NaCl 8000 8000
KCl 400 400
KH2 PO4 60 60
Glucose 1000 1000
Na2 HPO4 48 48
MgSO4 - 98
CaCl2 - 140
Na2 CO3 350 350

3. Results

3.1. Microstructure Characterization


The microstructure of the selected samples of pure magnesium compacts and details showing
characteristic features of the material are presented in Figure 2. Microstructural analysis revealed
that with increasing compaction pressure, the deformation of the powder particle increased and the
porosity decreased. This trend was apparent in all the prepared samples compacted at elevated and
room temperature.

(a) (b)

(c) (d)
Figure 2. Cont.

54
Metals 2017, 7, 461

(e) (f)

(g) (h)
Figure 2. Microstructure of selected samples: (a,b) cold compacted at 100 MPa; (c,d) cold compacted at
500 MPa; (e,f) hot pressed at 100 MPa, 400 ◦ C; (g,h) hot pressed at 500 MPa, 400 ◦ C; pores are indicated
with arrows.

A significant change in porosity is visible when cold pressed (Figure 2a,b) and hot pressed
materials (Figure 2c,d) prepared under compacting pressure of 100 MPa are compared. In cold
compacted material, the porosity is open, and metallographic resin used for the sample molding is
present in the microstructure (arrows in Figure 2a,b). The hot pressed sample contained also some
pores in the microstructure (Figure 2e); however, they were much smaller when compared to the cold
pressed sample (Figure 2a). Only the oxide layer present on the powder particles’ surface can be seen
in Figure 2f; no metallographic resin is observed in the pores. Hot pressing under 500 MPa at 400 ◦ C
led to a further decrease in porosity (Figure 2g,h).
Applying higher pressures during sample preparation led to a significant decrease of porosity
in the cold compacted samples, as well as in the hot pressed samples. The level of deformation of
the powder particles increased with increasing the compacting pressure. However, the change was
observed to be increasing from 100 MPa up to 300 MPa compacting pressures, while the shape and
the level of deformation of the powder particles in the case of cold and hot pressed samples did not
change for compacting pressures from 300 MPa to 500 MPa. The size of metallic grains within the
powder particles was impossible to calculate using light microscopy (Figure 2). Electron backscattered
diffraction was therefore applied to reveal the basic metallic structure of the samples (Figure 3).
The main difference in the microstructure between cold and hot pressed samples applying
higher pressures than 300 MPa was revealed by EBSD. In Figure 3, EBSD maps of cold compacted
and hot pressed (400 ◦ C, 1 h) samples prepared under 400 MPa are shown. In the case of the cold
compacted sample, very fine grains were created in the powder particle during the material compaction
(Figure 3a,b). However, the material on the powder particles’ interface was not identified by EBSD,
which indicates that the powder particles are not diffusion bonded. Larger grains were created in the
magnesium powder particles compacted at 400 ◦ C (Figure 3c,d). The observed metallic grains are
connected to the metallic grains created in the neighboring powder particles, which indicates the good
connection of the material structure and promotes powder particles’ bonding.

55
Metals 2017, 7, 461

(a) (b)

(c) (d)

Figure 3. Electron backscattered diffraction (EBSD) analysis of cold compacted and hot pressed material:
(a) SEM image and (b) EBSD map of the cold compacted sample prepared under 400 MPa; (c) SEM
image and (d) EBSD map of the hot pressed sample prepared under 400 MPa at 400 ◦ C. Colored-metallic
grains of magnesium: dark—grain boundary (lines), pores or oxides (wide gaps).

3.2. Mechanical Characterization

3.2.1. Three-Point Bending Test and Microhardness Test Results


Three-point bending test revealed an increasing trend of flexural strength of prepared materials
corresponding to the increasing samples’ compacting pressure (Figure 4a). The highest flexural
strength was obtained for hot pressed samples compacted at 400 ◦ C. Samples prepared at 300 ◦ C
followed the same trend as the samples prepared at 400 ◦ C; however, the flexural strength was lower
(Figure 4a). Hot pressing magnesium powder at 500 ◦ C resulted in lower values of flexural strength
compared to lower compaction temperatures (except RT (room temperature)). Furthermore, during the
preparation of the samples pressed under 400 MPa at 500 ◦ C, deformation of the steel mold (used for
sample preparation) occurred; therefore, the samples’ preparation was not possible. The cold pressed
samples’ test results show significantly lower values of flexural strength compared to the hot pressed
samples. Except the flexural strength measured on hot pressed samples prepared under 500 MPa,
the value of flexural strength was observed to increase with increasing compaction pressure.

56
Metals 2017, 7, 461

(a) (b)
Figure 4. Three-point bending tests and microhardness measurement results; (a) flexural strength, (b)
microhardness reuslts (RT: room temperature).

Microhardness test results revealed a significant difference between cold and hot pressed samples
(Figure 4b). The increase of the microhardness corresponding to the increasing compaction pressure was
observed for cold compacted and hot pressed samples prepared at 300 ◦ C and 400 ◦ C (Figure 4b), while the
identical trend of microhardness evolution was observed for both hot pressed materials. The influence of
compaction pressure on the microhardness was more pronounced for cold compacted samples, however,
higher values of microhardness were characteristic for hot pressed samples. A specific situation was
observed in the case of hot pressed samples prepared at 500 ◦ C, where the primary microhardness increase
observed when increasing compaction pressure from 100 MPa to 200 MPa was followed by a decrease of
the microhardness for samples prepared under 300 MPa (Figure 4b).

3.2.2. Fractographic Analysis


Fracture surfaces of samples broken during the three-point bending test were analyzed in the
region where tensile loading was applied. The fractographic analysis revealed differences between
cold compacted and hot pressed materials (Figure 5).

(a) (b)

(c) (d)
Figure 5. Selected fracture surfaces of samples’ failure during three-point bending test: (a) cold
compacted 100 MPa; (b) cold compacted 500 MPa; (c) hot pressed 100 MPa at 400 ◦ C; (d) hot pressed
500 MPa at 400 ◦ C.

57
Metals 2017, 7, 461

Cold compacted materials’ fracture surfaces exhibited only trans-granular failure regardless of
the applied compaction pressure. Different applied compacting pressure resulted only in differences of
deformation of the powder particles present on the fracture surface (Figure 5a,b). Samples compacted
under 100 MPa revealed only minimal plastic deformation of the powder particles, and the fracture
surface exhibited a quite large amount of secondary cracks following magnesium powder particles’
boundaries (Figure 5a). Samples compacted under higher pressures (300 MPa and higher) exhibited
significant plastic deformation of powder particles and a smaller amount of secondary cracks compared
to the low pressure compacted samples, as was observed on the samples’ fracture surfaces (Figure 5b).
Hot pressed samples contained significantly deformed powder particles, and only trans-granular
failure was observed for samples compacted under lower pressures (100 MPa and 200 MPa) (Figure 5c).
With the increase of the compaction pressure in the case of hot pressed samples, also the inter-granular
failure of the powder particles was observed on the samples’ fracture surfaces (Figure 5d, marked with
an arrow).

3.3. Electrochemical Characterization


The corrosion resistance of the studied materials was evaluated by EIS. As the corrosion medium,
HBSS and enriched HBSS (containing Mg2+ and Ca2+ ions) were used (the composition is given in
Table 1). Based on the obtained data represented by Nyquist plots, three equivalent circuits (EC)
shown in Figure 6 were used for the data analysis and determination of materials’ electrochemical
corrosion characteristics.

(a) (b)

(c)
Figure 6. (a–c) Equivalent circuits used for evaluation of the Nyquist plots.

The EC given in Figure 6a was used for the description of the behavior of materials exhibiting
two capacitive loops on the Nyquist plot. One capacitance loop was obtained for high frequencies and
the second capacitance loop for low frequencies. Two capacitance loops characterize the response of
the material by the creation of a layer of corrosion products on its surface. The EC in the Figure 6a
can be from the electrochemical point of view explained as the simple electrical response of the
damaged (porous and cracked) layer of corrosion products created on the metal surface to the corrosion
environment (partially blocked electrode [33]). Damage of the layer can be understood for example
as cracks in the layer created due to its porous nature and thickness increasing due to the material
response to the corrosion environment, while some parts of the damaged layer can be removed from
the material surface, and the metallic surface can be revealed to the corrosion environment again.
EC presented in Figure 6a consists of elements representing solution resistance R1 ; Q2 is a constant

58
Metals 2017, 7, 461

phase element (CPE), which is a component modeling the behavior of the present corrosion products
(taking into account its porous origin and its damage); R2 represents the resistance of the corrosion
products’ layer against the solution including the resistance of the corrosion products and resistance of
the material against the solution penetrating through the corrosion products’ layer damaged areas
and pores; Q3 is a constant phase element (CPE), which is a component modeling the behavior of
the interface between the solution and metal surface including the influence of the damaged layer of
corrosion products, and R3 is the resistance of the material against the solution (charge transfer in the
double layer).
The resulting polarization resistance Rp of this model can be calculated according to the Equation (1):

Rp = R2 + R3 (1)

EC including an inductive element is shown in Figure 6b. The model was used for the evaluation
of the Nyquist plots containing one high-frequency capacitance loop and one low-frequency inductive
loop. The element R1 represents solution resistance; Q2 represents double layer capacitance expressed
by the CPE element; R2 is the resistance against the charge transfer of the solution and the created layer
of corrosion products; L3 is inductance connected with the negative difference effect (NDE) caused
in the case of magnesium by adsorption of H+ , Mg+ [30]; and R3 is the resistance of this inductive
element and also the resistance of charge transfer on the sample surface. The resulting polarization
resistance Rp value of this model is given by Equation (2):

1 1 1
= + (2)
Rp R2 R3

In the case of Nyquist plots consisting of the high-frequency capacitance loop, medium frequency
capacitance loop and low-frequency inductive loop, an EC shown in the Figure 6c was used for the
data analysis. In this case, the creation of the layer of corrosion products (porous in its nature) on
the material surface accompanied with additional creation of the secondary layer (outer) of corrosion
products on its surface closing the pores in the primary (inner) layer is assumed. The element R1
represents the solution resistance; Q2 can be connected with the outer layer of corrosion products; R2 is
the resistance against charge transfer of the outer layer; Q3 represents the capacitance between the
outer and inner porous layer of corrosion products; R3 is the resistance against charge transfer between
the outer and inner porous layers of corrosion products; L4 is inductance equivalent to the response of
the adsorbate species on the sample surface due to the NDE; and R4 is resistance against the charge
transfer of interfacial reaction on the sample surface (resistance caused by adsorbed species on the
metal surface). The final Rp in the model is given by Equation (3):

R3 · R4
Rp = R2 + (3)
R3 + R4

3.3.1. EIS Measured in HBSS


The EIS spectra for cold compacted samples measured in HBSS medium in the form of Nyquist
plots are shown in Figure 7a, and data obtained by fitting the plots are given in Table 2. The plots
contain three loops including an inductive one. Based on the obtained plots’ character assuming
the same material response to HBSS in time, the EC model shown in Figure 6c was used for the
electrochemical corrosion characteristics’ determination. Measurement of the cold compacted sample
was stopped after 12 h due to the fatal corrosion damage of the samples. The very low capacitance
response between the layer of corrosion products and compacts was observed at the beginning of
the exposure of the samples to the corrosion environment. This could be explained by difficulties
in the creation of a coherent layer of corrosion products on the sample’s surface due to the high
porosity of the cold compacted material. The values of Rp were only slowly increasing up to the end

59
Metals 2017, 7, 461

of the EIS measurements. The initial Rp value was determined as 109 Ω·cm2 . An increase of Rp was
observed with increasing the immersion time to 2 h; however, this increase was followed by a drop
of Rp to a value comparable with the initial polarization resistance. This drop was followed by a
slow increase of the polarization resistance of cold compacted samples up to 252 Ω·cm2 obtained after
12 h. Changing values of n2 and n3 , characterizing the stability of the created outer and inner layer of
corrosion products on the sample surface indicate the non-uniformity of the layer. At the beginning
of the measurement, the inner layer of corrosion products exhibited higher stability, and its higher
contribution to the protectivity of the samples was changed in immersion times of 2 h and 4 h when
the outer layer with higher stability could protect the material against the following corrosion attack.
However, exceeding 4 h of exposure of the material to HBSS, the created layers exhibited only very
low stability (expressed by coefficients n2 and n3 ) (Table 2).

(a) (b)

(c) (d)
Figure 7. Electrochemical behavior of powder metallurgy (PM) magnesium materials in HBSS: (a) cold
pressed 500 MPa; (b) hot pressed 100 MPa at 400 ◦ C; (c) hot pressed 500 MPa at 400 ◦ C; (d) comparison
of Rp values in time.

Table 2. Fitted EIS (electrochemical impedance spectroscopy) data for the cold compacted sample
prepared under 500 MPa, HBSS.

R1 R2 R3 R4 Rp Q2 Q3
Time (h) n2 n3 L4 (H)
(Ω·cm2 ) (Ω·cm2 ) (Ω·cm2 ) (Ω·cm2 ) (Ω·cm2 ) (F·sn −1 ·10−6 ) (F·sn −1 ·10−6 )
0 65 130 64 −16 109 0.50 72.50 0.61 0.97 390
1 63 244 0 0 196 279.00 715.00 0.67 0.60 0
2 66 51 161 74 101 284 0.17 0.91 0.74 2
4 63 54 145 219 141 22.16 0.00 1.00 0.54 1
8 60 190 −1 −8 190 520.00 1.58 0.47 0.01 42
12 77 189 95 189 252 221 5.72 0.67 0.56 1

60
Metals 2017, 7, 461

The characteristic Nyquist plots for the hot pressed sample prepared under 100 MPa obtained
with EIS in HBSS are shown in Figure 7b, and the data obtained after fitting of these curves applying
adequate EC are shown in the Table 3. The initial and the last measured electrochemical response of
the sample tested in HBSS (measurements at 0 and 12 h) correspond to behavior that can be described
by the EC model given in Figure 6a. In other cases, three loops were characteristic for Nyquist
plots, and the compacts’ behavior can be described with EC in Figure 6c. This type of electrical
response includes also an inductive loop. The hot pressed 100 MPa powder compacts were measurable
only within 12 h due to the high degradation of the material in HBSS. The Rp values rose from the
beginning of the experiment (305 Ω·cm2 ) to the maximum value (627 Ω·cm2 ) obtained after 1 h of
exposure of the material to HBSS. After 1 h of exposure, the Rp continuously decreased until the
end of the experiment (251 Ω·cm2 ). The value of n2 characterizing the stability of the outer layer of
corrosion products decreased with increasing the time of exposure of the compacts to the corrosion
environment. Nevertheless, the value of n3 characterizing the inner porous layer of corrosion products
(created in the areas of contact of the corrosion solution with the damaged primary created layer
of corrosion products) was changing in the whole experiment time, indicating its reaction with the
compacted material. However, the primary created layer exhibited good stability and could exhibit the
protection of the material in the corrosion in times of 0 h, 4 h and 12 h of exposure (Table 3). However,
the decreasing n3 values indicate the primary layer damage. Changing values of n2 and n3 characterize
the stability of layers of corrosion products created on the material surface.

Table 3. Fitted EIS data for the hot pressed sample prepared under 100 MPa at 400 ◦ C, HBSS.

R1 R2 R3 R4 Rp Q2 Q3
Time (h) n2 n3 L4 (H)
(Ω·cm2 ) (Ω·cm2 ) (Ω·cm2 ) (Ω·cm2 ) (Ω·cm2 ) (F·sn −1 ·10−6 ) (F·sn −1 ·10−6 )
0 54 263 43 - 305 73.9 2676.0 0.83 1.00 -
1 50 627 −1 −1 627 132.1 8887.0 0.74 0.69 0
2 49 655 11 −10 527 136.4 7759.0 0.74 0.39 4
4 51 474 3969 0 474 176.8 6082.0 0.70 0.99 136
8 48 270 162 −2 268 206.9 5617.0 0.68 0.03 2
12 53 −3 254 - 251 262.6 26.6 0.55 0.86 -

The Nyquist curves for hot pressed samples prepared under 500 MPa are shown in Figure 7c, and the
data obtained after fitting of the curves are shown in Table 4. Only one type of electrochemical response
was identified for all the measurements of the samples. The suitable EC model for material behavior
analysis is shown in Figure 6b. The EC in Figure 6b is composed of two time-independent loops and
contains one inductive loop. The inductive electrical response of the material to the HBSS was clear for
the whole experiment time. The material polarization resistance Rp increased from the initial (396 Ω·cm2 )
measurement and reached the maximum value (892 Ω·cm2 ) at the measurement at 4 h of exposure.
This initial increase of Rp was followed by a gradual decrease until 72 h of exposure (246 Ω·cm2 ). The last
Rp value after 96 h of exposure was 261 Ω·cm2 . The value of n2 , in Table 4, indicates similar stability of
the created layer of corrosion products on the material surface in the whole exposure time.
The comparison of the evolution of Rp of the studied samples in HBSS in time is shown in
Figure 7d. In all the cases, an initial increase of Rp was observed; however, in the case of cold
compacted and hot pressed material prepared under 100 MPa, this increase was observed only until 1
h of exposure. Cold pressed samples reached lower values of polarization resistance when compared
to the hot pressed samples. The primary increase of Rp was followed by a decrease for 2 h of exposure,
and the following increase was observed for cold compacted samples. The value of polarization
resistance for hot pressed samples prepared under 100 MPa was after the maximum reached after 1
h of exposure only decreasing. A similar trend was characteristic also for the hot pressed samples
prepared under 500 MPa; however, in this case, the maximal value of Rp was reached after 4 h of
exposure. All the samples reached similar values of polarization resistance (before samples’ fatal
degradation) at the maximal time of exposure.

61
Metals 2017, 7, 461

Table 4. Fitted EIS data for the hot pressed sample prepared under 500 MPa at 400 ◦ C, HBSS.

R1 R2 R3 R4 Rp Q2 Q3
Time (h) n2 n3 L3 (H)
(Ω·cm2 ) (Ω·cm2 ) (Ω·cm2 ) (Ω·cm2 ) (Ω·cm2 ) (F·sn −1 ·10−6 ) (F·sn −1 ·10−6 )
0 54 289 107 - 396 62.1 - 0.79 - 1005
1 55 354.5 164 - 519 62.2 - 0.79 - 762
2 56 456.5 169 - 626 58.0 - 0.79 - 540
4 61 647 245 - 892 40.4 - 0.81 - 733
8 65 512 307 - 819 31.6 - 0.82 - 2067
12 66 487 283.5 - 771 30.3 - 0.81 - 2081
24 69 368 276 - 644 31.6 - 0.79 - 1725
48 74 164 156 - 320 48.9 - 0.76 - 796
72 72 114 132 - 246 69.0 - 0.74 - 703
96 78 130 131 - 261 69.0 - 0.75 - 744

3.3.2. EIS Measured in Enriched HBSS


The electrochemical corrosion tests were also performed on the prepared samples in the enriched
HBSS with Mg2+ and Ca2+ .
The EC in Figure 6b with three loops was selected for analysis of the Nyquist plots obtained for
cold pressed samples prepared under 500 MPa up to 12 h of exposure to enriched HBSS. The Nyquist
plot characterizing material behavior in 24 h of exposure to the corrosion environment had a different
character, and the EC for the partially blocked electrode shown in Figure 6a was used for the data
determination. After 24 h of exposure of the samples to the corrosion environment, substantial
corrosion damage occurred. The determined data are shown in the Table 5, and the obtained Nyquist
plots are shown in Figure 8a. The Rp obtained at the beginning of the measurement (473 Ω·cm2 )
was followed by its decrease to a value of 104 Ω·cm2 obtained after 8 h of exposure. The value of
polarization resistance started to increase again, and the value of 530 Ω·cm2 was determined at the end
of the experiment (24 h). The stability of the layer of corrosion products created on the samples’ surface
is characterized by the values of n2 and n3 . Based on the Nyquist plots’ character (EC in Figure 6b
and EC in Figure 6a), the stability of both layers is changing in time, which indicates their changing
reactivity and degradation in enriched HBSS in time.

(a) (b)

(c) (d)

Figure 8. Electrochemical behavior of PM magnesium materials in enriched HBSS: (a) cold pressed
500 MPa; (b) hot pressed 100 MPa 400 ◦ C; (c) hot pressed 500 MPa 400 ◦ C; (d) comparison of Rp values
in time.

62
Metals 2017, 7, 461

Table 5. Fitted EIS data for the cold compacted sample prepared under 500 MPa, enriched HBSS.

R1 R2 R3 R4 Rp Q2 Q3
Time (h) n2 n3 L (H)
(Ω·cm2 ) (Ω·cm2 ) (Ω·cm2 ) (Ω·cm2 ) (Ω·cm2 ) (F·sn −1 ·10−6 ) (F·sn −1 ·10−6 )
0 56 703 55 −45 473 63.6 126.3 0.72 0.10 71
1 67 327 376 204 459 105.3 95.9 0.71 0.47 0
2 60 208 129 −3 205 689.8 151.1 0.28 0.80 996
4 77 28 156 2866 176 24.3 285.4 0.65 0.63 0
8 76 147 6 −5 104 644.1 276.0 0.44 0.95 2
12 71 137 115 291 219 135.4 915.5 0.43 0.60 0
24 80 423 107 - 530 28.6 652.6 0.49 0.68 -

The Nyquist plots characterizing the behavior of hot pressed samples prepared under 100 MPa
at 400 ◦ C are presented in Figure 8b. Based on the character of the obtained plots, the EC given
in Figure 6b (up to 4 h of exposure) and Figure 6a (8 to 24 h of exposure) were used for the data
determination. The measurement was performed for 24 h because of the high level of sample corrosion
degradation after this time. The initial Rp value was determined as 83 Ω·cm2 by fitting analysis.
The initial state was followed by a slight increase and decrease of Rp values until the end of the
measurement when the maximal value of Rp was determined as 319 Ω·cm2 . The stability of the layer
of corrosion products (expressed with the n2 value given in Table 6) created on the sample’s surface
was decreasing with increasing exposure time; however, after 8 h of exposure, the layer became porous,
and some reactions took place also in the pores where the corrosion environment reached the surface
of the samples.

Table 6. Fitted EIS data for the hot pressed sample prepared under 100 MPa at 400 ◦ C, enriched HBSS.

R1 R2 R3 R4 Rp Q2 Q3
Time (h) n2 n3 L (H)
(Ω·cm2 ) (Ω·cm2 ) (Ω·cm2 ) (Ω·cm2 ) (Ω·cm2 ) (F·sn −1 ·10−6 ) (F·sn −1 ·10−6 )
0 50 243 125 - 83 63.9 - 0.75 - 2799
1 52 255 231 - 121 109.8 - 0.69 - 2368
2 52 240 360 - 144 220.2 - 0.60 - 2693
4 51 176 539 - 133 366.5 - 0.53 - 3512
8 56 122 41 - 163 191.6 91.4 0.58 0.92 -
12 54 25 123 - 147 72.4 190.1 0.61 0.65 -
24 70 219 100 - 319 63.9 171.1 0.41 0.77 -

The EC given in Figure 6a was used for the fitting of the Nyquist plots characterizing the behavior
of hot pressed samples prepared under 500 MPa at 400 ◦ C at the beginning of the exposure (0 h)
(Figure 8c). Increasing the exposure time, the character of the Nyquist plots changed, and the EC
shown in Figure 6b was used for the plots’ fitting. Except the first measurement, there is a clear
inductive response in all the Nyquist plots. The obtained electrochemical corrosion characteristics are
given in Table 7. The Rp increases from the initial (436 Ω·cm2 ) measurement and reached the maximum
value (647 Ω·cm2 ) at 12 h of measurement. This increase of Rp was followed by a continual decrease
until the end of the experiment at 96 h of exposure (321 Ω·cm2 ). The stability of the layer of corrosion
products created on the samples’ surface characterized with the n2 value slightly decreased after 2 h of
exposure to the enriched HBSS, however, remained stable, and its stability started to slightly increase
after 48 h of exposure (Table 7).
Rp evolution in enriched HBSS in time determined for magnesium PM processed samples is
shown in Figure 8d. The value of Rp determined for the cold compacts at the beginning of the exposure
decreased with time of exposure up to 8 h, while the following exposure resulted in the Rp increase
to a value slightly higher than the primary value. In the case of hot pressed samples prepared under
100 MPa at 400 ◦ C, a slight increase of the Rp from the beginning of the exposure up to 12 h was
observed, while additional exposure to the enriched HBSS was connected with the following increase
of Rp . In the case of hot pressed samples prepared under 500 MPa at 400 ◦ C, an increase of Rp up
to 12 h of exposure was observed, while additional exposure resulted in a polarization resistance

63
Metals 2017, 7, 461

decrease to values below the Rp measured at the beginning of the exposure. However, the values
of Rp determined for hot pressed samples prepared under 500 MPa were higher compared to other
material states.

Table 7. Fitted EIS data for the hot pressed sample prepared under 500 MPa at 400 ◦ C, enriched HBSS.

R1 R2 R3 R4 Rp Q2 Q3
Time (h) n2 n3 L (H)
(Ω·cm2 ) (Ω·cm2 ) (Ω·cm2 ) (Ω·cm2 ) (Ω·cm2 ) (F·sn −1 ·10−6 ) (F·sn −1 ·10−6 )
0 53 332 104 - 436 32.3 2832.0 0.85 0.49 -
1 64 262 215 - 476 45.2 - 0.74 - 1526
2 65 261 314 - 575 54.6 - 0.68 - 2367
4 67 302 326 - 629 53.8 - 0.66 - 2520
8 72 288 352 - 640 53.6 - 0.65 - 2268
12 72 260 387 - 647 53.5 - 0.65 - 3215
24 84 300 300 - 600 52.8 - 0.65 - 2246
48 83 194 226 - 419 57.0 - 0.71 - 2139
72 84 168 157 - 325 37.3 - 0.77 - 1410
96 77 156 165 - 321 42.4 - 0.76 - 1565

3.4. Corrosion Mechanism Analysis


The surface of the compacts tested by the EIS method was documented with the aim to compare
the mechanism of corrosion attack and its dependence on material structure and the corrosion
environment used (Figures 9 and 10). Metallographic cross-sections of the tested compacts were
prepared with the aim to identify the mechanism of the corrosion process within the material volume.
Samples’ cross-sections were molded into the resin and prepared by the standard metallographic
procedures. Due to the molding into the resin and the porosity and incompactness of the corroded
compacts and corrosion products on their surface, gas bubbles were present on the mold specimens
(above the surface of the sample).

(a) (b)

(c) (d)

Figure 9. Cont.

64
Metals 2017, 7, 461

(e) (f)

Figure 9. Corroded surface of PM magnesium compacts (the area exposed during EIS measurement is
outlined) and specimens cross-sections (A, resin; B, layer of corrosion products; C, compact); (a) cold
compacted under 500 MPa, 12 h in HBSS; (b) cold compacted under 500 MPa, 24 h in enriched HBSS;
(c) hot pressed under 100 MPa, 400 ◦ C, 12 h in HBSS; (d) hot pressed under 100 MPa, 400 ◦ C, 24 h in
enriched HBSS; (e) hot pressed under 500 MPa, 400 ◦ C, 96 h in HBSS; (f) hot pressed under 500 MPa,
400 ◦ C, 96 h in enriched HBSS.

(a) (b)

(c) (d)

(e) (f)

Figure 10. Corroded cross-section of pure magnesium compacts: (a) cold compacted 500 MPa, 12 h in
HBSS; (b) cold compacted 500 MPa, 24 h in enriched HBSS; (c) hot pressed 100 MPa, 400 ◦ C, 12 h in
HBSS; (d) hot pressed 100 MPa, 400 ◦ C, 24 h in enriched HBSS; (e) hot pressed 500 MPa, 400 ◦ C, 96 h in
HBSS; (f) hot pressed 500 MPa, 400 ◦ C, 96 h in enriched HBSS.

65
Metals 2017, 7, 461

In all cases, the area exposed to the corrosion environment during the EIS measurement was
covered with a layer of corrosion products. The created layer was not compact and contained a large
number of cracks, which were observed also in the layer of corrosion products.
The surface of the area of the compacts exposed to the HBSS and enriched HBSS was fully covered
with grey-/white-colored corrosion products during the EIS measurement (outlined areas). Based on
the surface observation, the corrosion mechanism can be considered as the same in the HBSS and
enriched HBSS, as the corrosion products’ structure is similar and no specific material behavior was
observed. The corrosion mechanism, however, was dependent on the sample preparation.
In the case of cold pressed samples, the created corrosion products covered the whole area
exposed to the corrosion environments (Figure 9a,b). On the details of the metallographic cross-section
given in Figure 9a,b, quite pronounced corrosion attack of the material can be seen. The layer
of corrosion products grows to a depth of approximately 4000 μm during the exposure to both
corrosion environments. Due to the large thickness of the created layer of corrosion products, its
cracking and decohesion with the cold compacted materials can be seen on samples’ cross-sections in
Figure 9a,b. In the details of Figure 10a,b, layer-like corrosion progression into the material volume can
be recognized. The material behavior was the same for HBSS and enriched HBSS. However, from the
comparison of Figure 10a,b, eventual removal of the layer of corrosion products from the surface of
the compact tested in enriched HBSS can be observed as the time of exposure increased. Therefore,
the more degradable influence of enriched HBSS on compacts compared to HBSS can be concluded for
cold compacted pure magnesium.
The hot pressed samples (Figure 9c–f) corroded uniformly throughout the whole tested area.
Samples prepared under 100 MPa compacting pressure at 400 ◦ C exhibited a uniform layer of corrosion
product after exposure in HBSS, with a slightly layer-like structure, which is more apparent from the
detail in Figure 10c. As the corrosion products’ layer grew into the material volume, the uniform layer
cracked, which in long-term exposure leads to very severe corrosion attack. Enriched HBSS seems
to have a similar influence on the hot pressed sample. The corrosion products’ layer observable
on the sample cross-section shown in Figure 9d was comparable to the one created in HBSS
(Figure 9c). Both layers had a comparable thickness (less than 1500 μm), and both of them exhibited
cracks. Furthermore, the mechanism of the corrosion attack progress through the material shown
in Figure 10c,d was similar. The corrosion seems to progress to the material volume layer by layer,
following layers of compacted powders (layer-like corrosion progression). The corrosion attack was
less pronounced when compared to the cold compacted materials.
The hot pressed samples prepared under 500 MPa at 400 ◦ C exhibit different behaviors in terms
of corrosion attack in HBSS and enriched HBSS compared to the cold pressed and hot pressed samples
prepared under 100 MPa. The surface of the hot pressed samples prepared under 500 MPa was
completely covered with corrosion products; however, the depth to which the samples degraded
is much smaller compared to the samples prepared under 100 MPa at 400 ◦ C and cold pressed
samples (Figure 9e,f). This shows much better coherence of the powder particles forming the samples.
The depth of corrosion attack is small (approximately 500 μm) comparing to the other material states.
However, in this case, different responses of compacts on HBSS and enriched HBSS were observed.
Even from the cross-sections in Figure 10e,f, the layer-like corrosion progression is visible only for the
sample exposed to HBSS. In the case of the exposure of the compacts to the enriched HBSS, the layer
of corrosion products was removed from the material surface during the measurement (Figure 10f),
which indicates low layer cohesion to the material and/or high volume expansion of the created
corrosion products.

4. Discussion
Microstructural, mechanical and electrochemical corrosion characteristics of pure magnesium
materials prepared by powder metallurgy were analyzed with the aim to identify the influence of
processing parameters. Sets of cold compacted and hot pressed materials (300 ◦ C, 400 ◦ C and 500 ◦ C)

66
Metals 2017, 7, 461

were prepared from magnesium powder under 100 MPa, 200 MPa, 300 MPa, 400 MPa and 500 MPa
compacting pressures. While the material has potential in biomedical applications, the electrochemical
corrosion characteristics of the processed materials were analyzed in HBSS and enriched HBSS using
electrochemical impedance spectroscopy, and the data were extended with the analysis of the corrosion
attack within the material volume.
Low compacting pressures led only to minimal plastic deformation of magnesium powder
particles during cold compaction or hot pressing (Figure 2). Increasing the compaction pressure
(above 300 MPa) at RT had an influence on the deformation of the powder particles while increased
deformability of particles corresponded with increasing compaction pressure. The influence was
also sufficient for changing the character of the porosity from open porosity to closed porosity and
decreased the number of pores on the samples’ cross-sections, which is in agreement with [11–13].
Hot pressing led to a significant decrease of the porosity of the compacts and also to a change in the
character of the porosity (Figure 2) [4,32]. The open porosity characteristic for cold pressed samples
was replaced with closed localized pores observed in hot pressed samples’ microstructure. This change
can be attributed to better plastic deformation of magnesium at a higher temperature, resulting in
higher plastic deformation of powder particles and closing the open porosity present on particles’
boundaries due to the shape factor (spherical particles cannot fill the space 100%). The change of
the magnesium deformability results from the hexagonal close-packed (HCP) crystallographic lattice
structure of the material. While only one slip system is active at RT, the increase in temperature led to
activation of more slip systems. Furthermore, the twinning mechanism is dominant for HCP lattice
structures instead of the slip mechanism below 200 ◦ C [34,35].
Besides the improvement of magnesium deformability, the elevated temperature also influenced
the evolution of grains in the compacted powder particles, which was revealed with EBSD analysis
shown in Figure 3. The cold compacted sample revealed a very fine microstructure within the
powder particle, but wide gaps between powder particles were observed. EBSD analysis of the hot
pressed sample revealed coarser metallic grains in powder particles; however, the gaps between
powder particles were not visible as diffusion bonding at higher temperatures took place between
powder particles.
A coarser metallic microstructure generally leads to a decrease in hardness and tensile strength
compared to the fine-grained material [4,5]. However, the observed fine microstructure corresponds
only to the powder particles, while the particles’ bonding was observed to be weak (also, an open
porosity was observed). Therefore, this fine microstructure obtained by cold compacting of magnesium
did not affect the mechanical properties of the samples in a positive way due to the low compactness
of particle boundaries.
Improvement of mechanical properties by hot pressing of pure magnesium is visible in
Figures 3 and 4. Hot pressed samples prepared at 300 ◦ C revealed the same trend in increasing
flexural strength with increasing compaction pressure as samples prepared at 400 ◦ C; however, the
values of flexural strength are slightly higher for the series of samples prepared at 400 ◦ C. This fact can
be attributed to a higher diffusion rate of magnesium at 400 ◦ C, which leads to better diffusion bonding
between powder particles and possibly more pronounced grain coarsening at higher temperature [11].
Samples prepared at 500 ◦ C reached much lower values of flexural strength compared to other hot
pressed samples (comparing samples prepared under the same pressures) and only slightly higher
than cold compacted samples (Figure 4a). At the temperature of 500 ◦ C, the diffusion of magnesium is
even higher than at 400 ◦ C, which should lead to better powder particle diffusion bonding; however,
the higher temperature (500 ◦ C) can lead to intensive coarsening of metallic grains created in the
powder particles. As a result, the grain coarsening could be a reason for the flexural strength decrease
(Figure 4a), considering good material compaction based on EBSD analysis.
Bonding of powder particles into the coherent sample is also apparent from the microhardness
test results. Cold compacted samples follow the same trend of increasing microhardness with
increasing compaction pressure as samples prepared via hot pressing; however, the obtained values are

67
Metals 2017, 7, 461

significantly lower. This can be attributed to the good cohesion of samples prepared via hot pressing.
Cold compacted samples hold together only due to the powder particles’ interlocking as a result of
plastic deformation, while the hot pressed samples are interlocked and diffusion bonded. The good
cohesion of powder particles in the case of hot pressed samples is also visible from the fractographic
analysis provided in Figure 5, as inter-granular failure was visible only on fracture surfaces of samples
prepared by hot pressing. Fractographic analysis from cold compacted samples revealed only the
trans-granular failure of samples, and no inter-granular failure was observed. While the character of
fracture surfaces differs, only brittle fracture was reported in all the measured samples. The character
of fracture surfaces furthermore confirms the assumption that the samples prepared by cold pressing
are not diffusion bonded.
Differences between cold compacted and hot pressed samples were also observed during EIS
measurement. Low compactness of cold compacted samples compared to the hot pressed samples
resulted in lower corrosion resistance of cold compacted samples. This was proven by the corrosion
rate and the depth to which corrosion progressed. It is clearly evident that by dissolving of magnesium
under H2 evolution, the corrosion environment is alkalized (increase pH value), and the corrosion
rate of magnesium slows down [30]. Corrosion products created on magnesium are usually based on
MgO and Mg(OH)2 . The layer of corrosion products, especially the Mg(OH)2 compound, is easily
disturbed and dissolved in environments containing Cl− . The layer offers only very low protection
to magnesium against corrosion especially in environments containing aggressive ions [36]. HBSS is
a more complex environment containing phosphates and hydrogen phosphates. More compounds
in the corrosion environment enable more chemical reactions between the metallic surface, corrosion
environment and corrosion products [30]. Products of these reactions can create a layer of corrosion
products (outer layer) over the created MgO and Mg(OH)2 inner primary layer created on the sample
surface. Both layers develop during the corrosion process and are porous by their nature or the damage
of the compact areas can occur. All the types of corrosion products are created during the material
exposure to the environment and grow on the material surface at the same time. These layers cannot
be exactly distinguished.
According to the Le Chatelier theory, HBSS should be more aggressive for magnesium than
enriched HBSS. Comparing the corrosion process character of examined magnesium materials
(Figures 9 and 10), this theory was not clearly proven. The similar thickness of the layers of corrosion
products (Figure 9), even the progress of corrosion attack within the material volume (Figure 10),
was observed considering the same processing parameters. Corrosion attack progress was the only
observed difference between the cold compacted and hot pressed samples; whereas the cold compacted
samples that were only mechanically-bonded exhibited a more pronounced degradation process when
compared to hot pressed samples. The higher compacting pressure used (500 MPa), resulting in even
better material compaction, had a positive influence also on material corrosion resistance.
The highest corrosion resistance was reached for hot pressed magnesium materials prepared under
500 MPa in both corrosion mediums. This was proven by higher polarization resistance and duration
of the corrosion experiment compared to other material states. The time of experiments indicates more
pronounced corrosion attack of samples in HBSS compared to enriched HBSS. This difference was not
observed for hot pressed samples prepared under 500 MPa. However, the values of Rp determined
for the materials tested in both solution types did not exhibit significant differences, and the higher
aggressiveness of HBSS compared to enriched HBSS was not proven with the results. Slightly higher
values of Rp reached for different material states in enriched HBSS when compared to HBSS did not
show a significant influence of corrosion environment on material properties. However, similar values
of Rp were reached at the end of EIS test for all the materials in each corrosion environment. Different
material types exhibit similar corrosion behavior regardless of the corrosion environment used.

68
Metals 2017, 7, 461

5. Conclusions
Powder metallurgy processed magnesium compacts prepared at room temperature and at 300 ◦ C,
400 ◦ C and 500 ◦ C under pressures in the range from 100 MPa to 500 MPa were analyzed in terms of the
evolution of the microstructure, mechanical properties and electrochemical corrosion characteristics.
Microstructural analysis revealed a positive influence of increasing pressure and temperature
on material compaction and porosity. The increase of the compacting pressure used led to increasing
of the powder particles’ deformation, which corresponds to the decrease of compacts’ porosity.
Elevating pressing temperature resulted in significant enhancement of powder particles’ deformability
and a decrease of compacts’ porosity. EBSD analysis showed fine-grained structure within cold
compacted powder particles; however, it also revealed that the powder particles were not bonded.
Elevated temperature resulted in grain coarsening; however, the powder particles were bonded.
Flexural strength was observed to be enhanced up to five times for the hot pressed samples
compared to the cold compacted samples. The samples prepared at 500 ◦ C reached similar
values of flexural strength as cold compacted samples. While only the inter-granular fracture
mechanism was observed for cold compacted and hot pressed samples prepared under lower pressures
(below 300 MPa), samples prepared at 400 ◦ C using higher compacting pressures exhibited also a
trans-granular fracture. Vickers microhardness of hot pressed samples was comparable for all the
preparation temperatures, and the values were higher than for cold compacted samples (up to 40%).
Electrochemical impedance spectroscopy measurements revealed a positive influence of the hot
pressing on the electrochemical corrosion behavior of PM magnesium. For both corrosion environments
used, the highest values of polarization resistance in time were determined for the hot pressed
samples prepared under 500 MPa. The corrosion mechanism was similar to the cold pressed and hot
pressed samples prepared under low pressures, where a layer of corrosion products (thicker for cold
compacted samples) was observed on the tested samples’ surface. A similar response of both sample
types was observed using HBSS and enriched HBSS (considering the sample preparation conditions).
Less pronounced corrosion attack was observed for hot pressed samples prepared under 500 MPa, and
also, a similar material response for both the corrosion environment types was observed.
Applying temperatures of 300 ◦ C and 400 ◦ C and high pressures (300 to 500 MPa) for magnesium
powder compaction seems to have a significantly positive influence on material bonding, mechanical
and electrochemical corrosion properties. A higher compaction temperature was shown to have a
detrimental effect on material compaction when using compacting pressure above 200 MPa.

Acknowledgments: This work was supported by Project No. LO1211, Materials Research Centre and by Project
No. LO1202, NETME (New Technologies for Mechanical Engineering) center plus, projects of Ministry of
Education, Youth and Sports of the Czech Republic under the “National sustainability program”.
Author Contributions: Pavel Doležal and Matěj Březina conceived and designed the experiments. Matěj Březina,
Michaela Krystýnová and Jozef Minda performed the experiments. Matěj Březina and Jozef Minda
analyzed the data. Stanislava Fintová, Petr Ptáček, Josef Zapletal and Jaromír Wasserbauer contributed
reagents/materials/analysis tools. Matěj Březina and Stanislava Fintová wrote the paper.
Conflicts of Interest: The authors declare no conflict of interest. The founding sponsors had no role in the design
of the study; in the collection, analyses or interpretation of data; in the writing of the manuscript; nor in the
decision to publish the results.

References
1. Kubásek, J.; Dvorský, D.; Čavojský, M.; Vojtěch, D.; Beronská, N.; Fousová, M. Superior properties of
Mg-4Y-3RE-Zr alloy prepared by powder metallurgy. J. Mater. Sci. Technol. 2017, 33, 652–660. [CrossRef]
2. Zhou, T.; Yang, M.; Zhou, Z.; Hu, J.; Chen, Z. Microstructure and mechanical properties of rapidly
solidified/powder metallurgy Mg-6Zn and Mg-6Zn-5Ca at room and elevated temperatures. J. Alloys Compd.
2013, 560, 161–166. [CrossRef]

69
Metals 2017, 7, 461

3. Yan, Y.; Cao, H.; Kang, Y.; Yu, K.; Xiao, T.; Luo, J.; Deng, Y.; Fang, H.; Xiong, H.; Dai, Y. Effects of Zn
concentration and heat treatment on the microstructure, mechanical properties and corrosion behavior of
as-extruded Mg-Zn alloys produced by powder metallurgy. J. Alloys Compd. 2017, 693, 1277–1289. [CrossRef]
4. Chang, I.; Zhao, Y. Advances in Powder Metallurgy: Properties, Processing and Applications; Woodhead
Publishing: Cambridge, UK, 2013; p. 90.
5. Suwas, S.; Gottstein, G.; Kumar, R. Evolution of crystallographic texture during equal channel angular
extrusion (ECAE) and its effects on secondary processing of magnesium. Mater. Sci. Eng. A 2007, 471.
[CrossRef]
6. Shen, J.; Imai, H.; Chen, B.; Ye, X.; Umeda, J.; Kondoh, K. Deformation mechanisms of pure Mg materials
fabricated by using pre-rolled powders. Mater. Sci. Eng. A 2016, 658, 309–320. [CrossRef]
7. Stulikova, I.; Smola, B.; Vlach, M.; Kudrnova, H.; Piesova, J. Influence of powder metallurgy route on
precipitation processes in MgTbNd alloy. Mater. Charact. 2016, 112, 149–154. [CrossRef]
8. Manne, B.; Bontha, S.; Ramesh, M.R.; Krishna, M.; Balla, V.K. Solid state amorphization of Mg-Zn-Ca system
via mechanical alloying and characterization. Adv. Powder Technol. 2017, 28, 223–229. [CrossRef]
9. Bornapour, M.; Celikin, M.; Cerruti, M.; Pekguleryuz, M. Magnesium implant alloy with low levels of
strontium and calcium: The third element effect and phase selection improve bio-corrosion resistance and
mechanical performance. Mater. Sci. Eng. C 2014, 35, 267–282. [CrossRef] [PubMed]
10. Nayak, S.; Bhushan, B.; Jayaganthan, R.; Gopinath, P.; Agarwal, R.D.; Lahiri, D. Strengthening of Mg based
alloy through grain refinement for orthopaedic application. J. Mech. Behav. Biomed. Mater. 2016, 59, 57–70.
[CrossRef] [PubMed]
11. Wang, X.; Hu, L.; Liu, K.; Zhang, Y. Grain growth kinetics of bulk AZ31 magnesium alloy by hot pressing.
J. Alloys Compd. 2012, 527, 193–196. [CrossRef]
12. Kayhan, S.M.; Tahmasebifar, A.; Koç, M.; Usta, Y.; Tezcaner, A.; Evis, Z. Experimental and numerical
investigations for mechanical and microstructural characterization of micro-manufactured AZ91D
magnesium alloy disks for biomedical applications. Mater. Des. 2016, 93, 397–408. [CrossRef]
13. Meng, F.; Rosalie, J.M.; Singh, A.; Tsuchiya, K. Precipitation behavior of an ultra-fine grained Mg-Zn alloy
processed by high-pressure torsion. Mater. Sci. Eng. A 2015, 644, 386–391. [CrossRef]
14. Reddy, T.H.; Pal, S.; Kumar, K.C.; Mohan, M.K.; Kokol, V. Finite element analysis for mechanical response of
magnesium foams with regular structure obtained by powder metallurgy method. Procedia Eng. 2016, 149,
425–430. [CrossRef]
15. Vahid, A.; Hodgson, P.; Li, Y. New porous Mg composites for bone implants. J. Alloys Compd. 2017, 724,
176–186. [CrossRef]
16. Zhang, S.; Zheng, Y.; Zhang, L.; Bi, Y.; Li, J.; Liu, J.; Yu, Y.; Guo, H.; Li, Y. In vitro and in vivo corrosion and
histocompatibility of pure Mg and a Mg-6Zn alloy as urinary implants in rat model. Mater. Sci. Eng. C 2016,
68, 414–422. [CrossRef] [PubMed]
17. Khalajabadi, S.Z.; Abdul Kadir, M.R.; Izman, S.; Marvibaigi, M. The effect of MgO on the biodegradation,
physical properties and biocompatibility of a Mg/HA/MgO nanocomposite manufactured by powder
metallurgy method. J. Alloys Compd. 2016, 655, 266–280. [CrossRef]
18. Bornapour, M.; Muja, N.; Shum-Tim, D.; Cerruti, M.; Pekguleryuz, M. Biocompatibility and biodegradability
of Mg-Sr alloys: The formation of Sr-substituted hydroxyapatite. Acta Biomater. 2013, 9, 5319–5330. [CrossRef]
[PubMed]
19. Singh Raman, R.K.; Jafari, S.; Harandi, S.E. Corrosion fatigue fracture of magnesium alloys in bioimplant
applications: A review. Eng. Fract. Mech. 2015, 137, 97–108. [CrossRef]
20. Yazdimamaghani, M.; Razavi, M.; Vashaee, D.; Moharamzadeh, K.; Boccaccini, A.R.; Tayebi, L. Porous
magnesium-based scaffolds for tissue engineering. Mater. Sci. Eng. C 2017, 71, 1253–1266. [CrossRef]
[PubMed]
21. Drábiková, J.; Pastorek, F.; Fintová, S.; Doležal, P.; Wasserbauer, J. Improvement of bio-compatible AZ61
magnesium alloy corrosion resistance by fluoride conversion coating. Koroze Ochrana Mater. 2016, 60.
[CrossRef]
22. Shadanbaz, S.; Dias, G.J. Calcium phosphate coatings on magnesium alloys for biomedical applications:
A review. Acta Biomater. 2012, 8, 20–30. [CrossRef] [PubMed]
23. Chiu, K.Y.; Wong, M.H.; Cheng, F.T.; Man, H.C. Characterization and corrosion studies of fluoride conversion
coating on degradable Mg implants. Surf. Coat. Technol. 2007, 202, 590–598. [CrossRef]

70
Metals 2017, 7, 461

24. Carboneras, M.; Hernández, L.S.; Del Valle, J.A.; García-Alonso, M.C.; Escudero, M.L. Corrosion protection
of different environmentally friendly coatings on powder metallurgy magnesium. J. Alloys Compd. 2010, 496,
442–448. [CrossRef]
25. Jafari, S.; Singh Raman, R.K.; Davies, C.H.J. Corrosion fatigue of a magnesium alloy in modified simulated
body fluid. Eng. Fract. Mech. 2015, 137, 2–11. [CrossRef]
26. Tahmasebifar, A.; Kayhan, S.M.; Evis, Z.; Tezcaner, A.; Çinici, H.; Koç, M. Mechanical, electrochemical and
biocompatibility evaluation of AZ91D magnesium alloy as a biomaterial. J. Alloys Compd. 2016, 687, 906–919.
[CrossRef]
27. King, A.D.; Birbilis, N.; Scully, J.R. Accurate electrochemical measurement of magnesium corrosion rates;
a combined impedance, mass-loss and hydrogen collection study. Electrochim. Acta 2014, 121, 394–406.
[CrossRef]
28. Bender, S.; Goellner, J.; Heyn, A.; Schmigalla, S. A new theory for the negative difference effect in magnesium
corrosion. Mater. Corros. 2011. [CrossRef]
29. Liao, J.; Hotta, M.; Mori, Y. Improved corrosion resistance of a high-strength Mg-Al-Mn-Ca magnesium alloy
made by rapid solidification powder metallurgy. Mater. Sci. Eng. A 2012, 544, 10–20. [CrossRef]
30. Song, G.-L. Corrosion of magnesium alloys. In Woodhead Publishing in Materials; Woodhead Publishing:
Philadelphia, PA, USA, 2011; p. 640.
31. Čapek, J.; Vojtěch, D. Properties of porous magnesium prepared by powder metallurgy. Mater. Sci. Eng. C
2013, 33, 564–569. [CrossRef] [PubMed]
32. Čapek, J.; Vojtěch, D. Effect of sintering conditions on the microstructural and mechanical characteristics of
porous magnesium materials prepared by powder metallurgy. Mater. Sci. Eng. C 2014, 35, 21–28. [CrossRef]
[PubMed]
33. Orazem, M.E.; Tribollet, B. Electrochemical Impedance Spectroscopy; Wiley: Hoboken, NJ, USA, 2008; p. 560.
34. Drápala, J. Hořčík, Jeho Slitiny a Binární Systémy Hořčík—Příměs: Magnesium, Its Alloya and Mg-Admixture
Binary Systems; Vysoká škola báňská—Technická Univerzita: Ostrava, Czech Republic, 2004; p. 172.
35. Kang, F.; Li, Z.; Wang, J.T.; Cheng, P.; Wu, H.Y. The activation of c + a non-basal slip in magnesium alloys.
J. Mater. Sci. 2012, 47, 7854–7859. [CrossRef]
36. Xin, Y.; Huo, K.; Tao, H.; Tang, G.; Chu, P.K. Influence of aggressive ions on the degradation behavior of
biomedical magnesium alloy in physiological environment. Acta Biomater. 2008, 4, 2008–2015. [CrossRef]
[PubMed]

© 2017 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

71
metals
Article
Development of a Novel Degradation-Controlled
Magnesium-Based Regeneration Membrane for
Future Guided Bone Regeneration (GBR) Therapy
Da-Jun Lin 1 , Fei-Yi Hung 1, *, Hung-Pang Lee 2 and Ming-Long Yeh 2, *
1 Department of Materials Science and Engineering, National Cheng Kung University, Tainan 701, Taiwan;
larrylin111@hotmail.com
2 Department of Biomedical Engineering, National Cheng Kung University, Tainan 701, Taiwan;
qer6322129@gmail.com
* Correspondence: fyhung@mail.ncku.edu.tw (F.-Y.H.); mlyeh@mail.ncku.edu.tw (M.-L.Y.);
Tel.: +886-06-2757575 (ext. 62950) (F.-Y.H.); +886-06-2757575 (ext. 63429) (M.-L.Y.)

Received: 21 September 2017; Accepted: 3 November 2017; Published: 6 November 2017

Abstract: This study aimed to develop and evaluate the ECO-friendly Mg-5Zn-0.5Zr (ECO505) alloy
for application in dental-guided bone regeneration (GBR). The microstructure and surface properties
of biomedical Mg materials greatly influence anti-corrosion performance and biocompatibility.
Accordingly, for the purpose of microstructure and surface modification, heat treatments and surface
coatings were chosen to provide varied functional characteristics. We developed and integrated both
an optimized solution heat-treatment condition and surface fluoride coating technique to fabricate a
Mg-based regeneration membrane. The heat-treated Mg regeneration membrane (ARRm-H380) and
duplex-treated regeneration membrane group (ARRm-H380-F24 h) were thoroughly investigated
to characterize the mechanical properties, as well as the in vitro corrosion and in vivo degradation
behaviors. Significant enhancement in ductility and corrosion resistance for the ARRm-H380 was
obtained through the optimized solid-solution heat treatment; meanwhile, the corrosion resistance of
ARRm-H380-F24 h showed further improvement, resulting in superior substrate integrity. In addition,
the ARRm-H380 provided the proper amount of Mg-ion concentration to accelerate bone growth
in the early stage (more than 80% new bone formation). From a specific biomedical application
point of view, these research results point out a successful manufacturing route and suggest that the
heat treatment and duplex treatment could be employed to offer custom functional regeneration
membranes for different clinical patients.

Keywords: Mg alloy; regeneration membrane; guided bone regeneration; heat treatment; fluoride
coating; biocompatibility

1. Introduction
Periodontitis is a bacterial-mediated inflammatory disease that can lead to damage of the
periodontal ligament and gingival tissue, and may also cause alveolar bone resorption [1].
The prevalence of continuous periodontitis growth has been well documented in modern society [2],
and in many clinical reports, delaying therapy can eventually result in tooth loss and alveolar atrophy
(especially for older patients) [3]. In such cases, the defect area must first be reconstructed before
artificial tooth root implantation. To this end, guided bone regeneration (GBR) procedures have been
noted as a reliable periodontal regeneration and alveolar augmentation therapy, and have registered
high success rates in recent years [4]. Currently, there are two material systems for GBR procedures,
namely degradable and non-degradable membrane materials [5]. With degradable regeneration
membranes (usually made by poly-lactic acid (PLA) or collagen sheets), secondary surgery for implant

Metals 2017, 7, 481; doi:10.3390/met7110481 72 www.mdpi.com/journal/metals


Metals 2017, 7, 481

removal is not required; however, the low mechanical strength and stiffness do not offer sufficient
structural strength. In contrast, non-degradable regeneration membranes (e.g., Ti mesh or Teflon-mesh)
are the most commonly used materials; however, their use requires secondary surgery for membrane
removal, which increases the risk of bacterial infection. In response to these clinical considerations, this
study targeted the development of a new type of regeneration membrane that features both adequate
strength and biodegradability.
Magnesium (Mg) is a promising metallic material for biomedical applications due to its unique
biodegradability, satisfactory biocompatibility, and excellent biomechanical properties [6]. Moreover,
Mg materials possess satisfactory biocompatibility and biofunctionality that can accelerate cell
proliferation and wound healing [7,8]. Therefore, Mg materials can be seen as potential candidates for
new types of regeneration membranes for dental GBR procedures. However, the poor anti-corrosion
behaviors and rapid mechanical fading of Mg materials in a physiological electrolyte environment
currently limit its clinical applicability [9,10]. Nevertheless, Cai et al. recently revealed that
Mg-5Zn alloy is a good candidate for orthopedic implants with an optimal Zn alloying amount [11].
Song et al. also demonstrated that Mg-5Zn alloy possesses a uniform corrosion behavior and less
localized corrosion incidence [12]. Moreover, several related studies have shown that properly
modified (including microstructure modification and surface treatment) Mg implants offer stable
mechanical retention with reasonable degradation rates both in vitro and in vivo [13–15]. Results
from our previous works showed that heat treatment and plastic deformation procedures greatly
influence the degradation behavior and biocompatibility of Mg-Zn-Zr series alloys via microstructure
transformation [7,16]. Li et al. investigated in vitro and in vivo corrosion, as well as the mechanical
properties and biocompatibility of Mg-Zn-Zr alloy; in addition, their research also indicated that the
degradation behavior and bone healing behavior could be effectively improved after surface fluoride
coating [17].
At present, the global demand of Mg materials is increasing year by year. It therefore follows
that the demand will further increase once Mg-based medical devices are widely accepted in
clinical practice. However, the casting procedure of Mg alloys always needs to use a great amount
of SF6 (a potent greenhouse gas), which can accelerate global warming [18]. To mitigate this
shortcoming, the SF6 -applied casting procedure should be replaced by novel green ECO-casting
techniques (using HFC-134a gas). In this study, to fabricate a highly functional medical device, a novel
ECO-casting technique, which integrates a solid-solution heat treatment and surface treatment, was
developed to produce an optimized biodegradable regeneration membrane for dental application
using ECO-friendly Mg-5Zn-0.5Zr (named ECO505) alloy material.

2. Materials and Methods

2.1. ECO-Casting Process and Shaping Process


Mg-20 wt. % Zr, 4N grade (99.99 wt. %) pure Mg and pure Zn ingots were used for producing a
Mg-5Zn-0.45Zr alloy billet. The precisely weighted raw materials were melted in a mild steel crucible
using an electrical resistance furnace. During the melting and alloying process, a 20% HFC-R134a-80%
Ar gas mixture was applied to the melt as a protection gas to prevent self-ignition [19]. The melt was
held at 700 ◦ C and stirred to homogenize the melt composition. After the melting process, the melt was
cast into a preheated (350 ◦ C) stainless steel mold to produce the ECO505 alloy billet (as shown in Figure 1).
Before the regeneration membrane shaping process, the cast billets were previously homogenized
at 380 ◦ C for 12 h and subsequently water quenched. Afterwards, the homogenized billets were
directly extruded at 350 ◦ C using a ram speed of 10 mm/s and an extrusion ratio of 35 to form stripe
extrudates. The rolling raw materials (40 mm in length, 15 mm in width and 4 mm in height) were
cut from the stripe extrudates with the rolling plane parallel to the extrusion direction. During the
hot-rolling process, the process conditions were fixed at the rolling strain rate of 1.98 s−1 and under
350 ◦ C, for which the reduction rate per rolling pass was 20%. Furthermore, the rolled samples were

73
Metals 2017, 7, 481

reheated to 350 ◦ C and kept at that temperature for 5 min prior to each rolling pass. The final thickness
of the as-rolled ECO505 regeneration membrane (named as ARRm) samples was controlled at 0.4 mm
(as shown in Figure 1).

Figure 1. Schematic diagram of ECO-casting process and shaping process.

2.2. Heat Treatment Modification


The ARRm samples were processed into a circular thin foil (12.7 mm in diameter). Solution
heat treatment of the ARRm samples was carried out in a tubular vacuum furnace at 380 ◦ C for
1, 2, 4 and 10 h, respectively. After heat treatment, the samples were water quenched at room
temperature. The names for the solution heat-treated samples were assigned by heating time as
follows: ARRm-380 ◦ C_1 h, ARRm-380 ◦ C_2 h, ARRm-380 ◦ C_4 h and ARRm-380 ◦ C_10 h. The name
of the final optimized condition was replaced by ARRm-H380 to show its representative meaning.

2.3. Surface Fluoride Coating


The surfaces of the ARRm and ARRm-H380 samples were coated with a protective MgF2 layer.
Prior to the coating process, all samples were polished successively with 200 to 8000 grit using
SiC sandpaper. The polished specimens were rinsed in acetone and ethanol, and subsequently
dried in a stream of dry air. The fluoride conversion treatment followed the suggested procedure
from a previously published report, which involved the samples being soaked in 42 wt. % HF and
placed in an orbital shaker at 90 rpm for 24 h [20]. After the samples were removed from the HF
conversion bath, they were rinsed twice with absolute ethanol and dried. Names for the MgF2 -coated
samples were assigned according to treatment history, and hereafter referred to as ARRm-F24 h and
ARRm-H380-F24 h, respectively.

2.4. Materials Characterization


All samples (as-cast, as-extruded and ARRm) for microstructure characterization were ground
and polished to 0.05 μm and finally etched by picric-acetic acid solution (4.2 g picric acid + 20 mL acetic
acid + 80 mL ethanol). Optical microscopy (BX41-LED, Olympus, Tokyo, Japan) and scanning electron
microscopy (JSM-6510LV, JEOL, Tokyo, Japan) with an energy dispersive spectrometer (EDS, INCA
350, Oxford Instrument, Oxford, UK) were used for microstructure characterization. The average grain

74
Metals 2017, 7, 481

size was determined according to the ASTM E112-96 standard. With the aim to characterize the effects
of heat treatment on matrix softening, Vicker’s hardness measurements (HMV-G21, Shimadzu, Kyoto,
Japan) were performed on the polished surface of the heat treated specimens with different treatment
conditions, for which the loaded force was 100 g and the holding time was 10 s. Tensile test used a
hydraulic-powered mechanical testing system (MTS-810, MTS Systems Corporation, Minneapolis,
MN, USA) with a tensile speed of 1 mm/min. A typical dog-bone tensile specimen expressed with a
gauge length of 20 mm, a gauge width of 7 mm and a thickness of 0.4 mm was selected for tensile test.
At least four samples are tested (Vicker’s hardness and tensile test) for each group.

2.5. In Vitro Corrosion Test


The electrochemical corrosion mechanism and performance of the fluoride conversion coating
were investigated using an electrochemical station (PARSTAT 2273, Princeton Applied Research,
Oak Ridge, TN, USA). Polarization curves were measured using a classic three-electrode cell,
where a saturated calomel electrode (SCE, +0.242 V vs. SHE) constituted the reference electrode
with a Pt-coated Ti mesh as the counter-electrode. The area of the working electrode exposed
to the electrolyte was controlled to within 1 cm2 by a Teflon holder. The electrolyte used was
revised simulated body fluid (r-SBF) solution (which per liter included 5.403 g of NaCl, 0.736 g
of NaHCO3 , 2.036 g of Na2 CO3 , 0.225 g of KCl, 0.182 g of K2 HPO4 , 0.310 g of MgCl2 ·6H2 O, 11.928 g of
4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid (HEPES), 0.293 g of CaCl2 , and 0.072 g of Na2 SO4
dissolved in deionized water) buffered at pH = 7.4 using HEPES and NaOH, with the environmental
temperature controlled at 37 ◦ C. Prior to conducting the experiment, nitrogen gas was bubbled through
the r-SBF to remove dissolved oxygen. The polarization curve was acquired with a scanning rate of
1 mV s−1 from −1.8 V to −0.8 V.
To characterize the corrosion properties of the Mg-based regeneration membrane, two immersion
corrosion examinations were used in this research. Firstly, the normal immersion corrosion
(un-bent/r-SBF immersion) test was employed to measure the corrosion rates, behaviors and tendencies
of the specimens in r-SBF by following the principle of ASTM G31-72. The r-SBF volume to surface area
ratio was fixed at 20 mL/cm2 . Secondly, the actual-simulated immersion corrosion (pre-bent/r-SBF
immersion) test was employed to acquire the corrosion rates, behaviors and tendencies of the pre-bent
specimens in r-SBF by a custom designed examination. For this examination, several PVC racks
(1 cm width and 5 cm long) were used to act as alveolar bone, upon which the pre-bent specimens
were fixed by nylon screws, as shown in Figure 2a. Then, the specimens were immersed into r-SBF.
The r-SBF volume to surface area ratio of this custom designed examination was fixed at 50 mL/cm2 .
The experiments were held for 1 week in an isothermal incubator at 37 ◦ C. The result is the average of
five samples (n = 5).

Figure 2. (a) Custom designed pre-bent/corrosion device and specimens; (b) animal
implantation model.

75
Metals 2017, 7, 481

2.6. Animal Model and Cranial Implantation Experiments


In this study, the animal experiments and surgical procedure were approved by the Institutional
Animal Care and Use Committee of National Cheng Kung University (approval No. 105258).
Male-controlled Sprague Dawley (SD) rats were housed in an environmentally controlled animal
feeding room (25 ± 1 ◦ C, 40~70% humidity, with a regular 12 h light cycle per day) to an age of
12~14 weeks. All SD rats were randomly assigned to testing groups prior to surgery. General anesthesia
was administered via an intra-abdominal injection of Zoletil 50 (Virbac, Carros, France) and Xylazine
(Panion and BF Biotech Inc., Taipei, Taiwan) mixture (0.2 mg/100 g of Zoletil 50 and 0.5 mg/100 g
of Xylazine); then, local infiltration anesthesia (xylocaine, 0.2 mL for each surgical incision) was
performed at the surgical sites before surgery. An incision along the periphery of the skull was created
to peel back to the anterior portion of the skull. Two 5 mm critical size defect (CSD) were drilled in the
calvarial bone using a trephine bur and low speed handpiece (as shown in Figure 2b). While drilling
the cranial bone, sterilized saline was continuously injected to cool the drilling heat. The defects were
covered with the regeneration membrane (a circular foil 7 mm in diameter and 0.4 mm thick) made by
ECO505 raw material. Control and experimental groups (ARRm-H380 and ARRm-H380-F24 h) were
employed to realize the applicability and performance of the regeneration membranes. After surgery,
the wounds were closed by 5–0 non-absorbable silk sutures and antibiotics applied to prevent wound
infection. At least four SD rats are examined for each group and time point (n = 4).

2.7. Micro-CT Analysis and 3D Image Reconstruction


To obtain both qualitative and quantitative data of the bone regeneration level within the CSD,
a μ-CT (Skyscan 1076, Kontich, Belgium) was used to scan the cranial bone. The sham, ARRm-H380
and ARRm-H380-F24 h groups were measured at 4-week and 12-week time points. The voltage and
X-ray current were controlled at 90 kV and 110 μA, respectively. The examined cranial bones and
ECO505 regeneration membranes were scanned through a 360◦ rotation angle, with a rotation interval
of 1◦ and pixel size of 18 μm resolution. From the data measurement with SkyScan software (Version
1.4.4, Kontich, Belgium), a cylindrical region of interest (ROI) 5 mm in diameter within the CSD site
was selected for analysis. The volume and diameter of the bone growth were measured as the new
bone volume fraction (%). The scanned cranial bone data were then reconstructed by Mimics software
(Version 4.0, Materialise NV, Leuven, Belgium) to obtain high-quality 3D reconstructions. For clear
identification, the original cranial bone and new bone were colored gray and light blue, respectively.

2.8. Statistical Analysis


The results of mechanical and in vivo implantation experiments are given as the mean
viability value ± standard deviation. Analysis of one-way variance (ANOVA) was conducted to
evaluate the statistical significance of differences. Differences at p ≤ 0.05 were considered to be
statistically significant.

3. Results and Discussion

3.1. Metallographic and Microstructure Observation


Figure 3 presents the optical micrographs of the as-cast, as-homogenized, as-extruded and
as-rolled specimens. For the ECO505 billet, abundant MgZn2 particles (the main secondary phase in
Mg-Zn-Zr alloy) distributed along the grain boundary can be seen. These brittle secondary phases
might affect the hot-working behavior of the Mg alloy. Prasad et al. suggested that applying
a homogenization heat treatment can dissolve the intermetallic particles and improve the hot
workability [21]. In the present study, after a homogenization treatment at 380 ◦ C for 12 h, most of the
MgZn2 phases were dissolved into the Mg matrix. The microstructure of the homogenized specimen
was characterized as equiaxed grains with an average grain size of about 90 μm. Figure 3d shows a
typical extrusion feature, in which dynamic recrystallization (DRX) grains, with an average grain-size

76
Metals 2017, 7, 481

of about 15 μm, can be observed in the matrix. Figure 3e,f show macrograph and optical images of the
ARRm. With respect to the macrograph, the ARRm showed no cracks, voids or severe-edge cracking.
Its microstructure was fully evolved to a fine DRX microstructure without twins or other harmful
defects, the average grain size of which was around 4.8 μm. The alloying elemental compositions of
the ARRm are presented in Table 1. As seen, ARRm contains low amounts of noble impurities (such as
Fe, Cr, Ni, Cu) and shows an acceptable level, indicating that the metallurgy and plastic deformation
process in this work are feasible.

Figure 3. Typical macrograph and/or microstructures of each group: (a) macrograph of ECO505
billet; (b) microstructure of as-cast ECO505; (c) microstructure of as-homogenized ECO505 billet;
(d) microstructure of as-extruded ECO505; (e,f) macrograph and microstructure of as-rolled ECO505
(named as ARRm).

Table 1. Elemental composition of ECO505 made regeneration membrane (unit: wt. %).

Group Zn Zr Fe Cr Ni Mn Cu Mg
ECO505 5.02 0.48 0.0012 0.0010 0.0023 0.0102 N.D. Bal.

77
Metals 2017, 7, 481

3.2. Solid-Solution Heat Treatment Modification and Mechanical Properties


Figure 4 shows the hardness variation curves of the ARRm samples isothermally heat treated at
340, 360, 380, 400 and 420 ◦ C for 1 h. For the specimens heat treated at 360 ◦ C, the hardness value
slightly decreased to 85 HV, which indicates insufficient solid solution efficiency. The hardness value
decreased with increasing heat treatment temperature, particularly for specimens heat treated at
temperatures higher than 380 ◦ C. For the treatment temperature of 380 ◦ C, the matrix was significantly
softened (65 HV) than 340 ◦ C and 360 ◦ C treated groups. This phenomenon also means that the brittle
secondary phases can be easily dissolved at 380 ◦ C; therefore, we selected the temperature control of
380 ◦ C for the ARRm heat-treatment process.

Figure 4. Vicker’s hardness variation curves of different heat treatment temperature. (Data presented
as mean ± SD, n = 4 and analyzed using a one-way ANOVA, * p < 0.05) (red dash line: the Vicker’s
hardness of un-solution heat treated specimen).

Figure 5 shows the microstructure evolution of ARRm isothermally heat treated at 380 ◦ C for 1 h,
2 h, 4 h and 10 h, the characteristics of which were used to determine the optimum heat treatment
time. Both the ARRm-380 ◦ C_1 h and ARRm-380 ◦ C_2 h featured many fine recrystallized grains,
revealing that the fully recovered matrices had equiaxed grains with an average grain size of 5.6 μm
and 7.8 μm, respectively. The average grain sizes of ARRm-380 ◦ C_4 h and ARRm-380 ◦ C_10 h
significantly grew to 18 μm and 26 μm, respectively. Notably, the tensile properties of ARRm-380 ◦ C_1 h
and ARRm-380 ◦ C_2 h showed an obvious improvement in elongation (see Figure 5e and Table 2).
According to many previous reports, this improvement is related to the grain boundary sliding
(GBS) phenomenon. GBS always occurs in fine-grained microstructures with a grain size smaller
than 10 μm [22]. With further increases in treatment time, the time variation significantly altered
the microstructure and tensile elongation. The tensile results clearly indicate that the mechanical
behavior of ARRm improved with increases in the 380 ◦ C heating duration until 2 h. Prolonging the
heating duration of the ARRm heat treatment beyond 4 h led to a serious drop in elongation and yield
strength behavior. Obviously, the specimens heat treated at 380 ◦ C for 4 h and 10 h had lower yield
strengths (189 MPa and 179 MPa, respectively) and elongations (11.2% and 11.1%, respectively) and
displayed the overheating condition, as evidenced by their big grain sizes not being able to trigger
GBS. Considering the application requirements, the elongation of the regeneration membrane must be
optimized for the clinical pre-bending procedure. In the present cases, the ARRm-380 ◦ C_2 h specimen
not only maintained a small grain size, but the brittle MgZn2 particles also dissolved, yielding a
synergistic effect that provided the best elongation of 20.2%. Therefore, the heat-treatment parameter
of 380 ◦ C for 2 h can be considered as the optimal condition for ARRm, and was named ARRm-H380
to show its representative meaning.

78
Metals 2017, 7, 481

Table 2. Mechanical parameter results of the tensile test (Data presented as mean ± SD, n = 4).

Parameter ARRm ARRm-380 ◦ C_1 h ARRm-380 ◦ C_2 h ARRm-380 ◦ C_4 h ARRm-380 ◦ C_10 h
UTS (MPa) 268 ± 12 251 ± 9 256 ± 5 224 ± 8 210 ± 5
YS (MPa) 248 ± 8 207 ± 7 200 ± 6 189 ± 6 179 ± 3
EL. (%) 8.5 ± 0.4 19.1 ± 0.6 20.2 ± 0.4 11.2 ± 0.5 11.1 ± 0.3

Figure 5. Typical microstructures and tensile stress-strain curves of each groups: (a) ARRm-380 ◦ C_1 h;
(b) ARRm-380 ◦ C_2 h; (c) ARRm-380 ◦ C_4 h; (d) ARRm-380 ◦ C_10 h, (e) stress-strain curves

3.3. Effect of Solid-Solution Treatment and Fluoride Coating on Anti-Corrosion Ability of Mg


Regeneration Membrane
Figure 6a,b show the surface morphologies of the fluoride coatings deposited on the ARRm
and ARRm-H380 substrate (named ARRm-F24 h and ARRm-H380-F24 h, respectively). The coatings
on both surfaces were neat and clean, without obvious coating defects or particle contaminants.
The growth reactions of the fluoride conversion coating on the magnesium surface are described in
the following:
Mg → Mg2+ + 2e− (1)

2H2 O + 2e− → H2 + 2OH− (2)

Mg2+ + 2OH− → Mg(OH)2 ↓, ΔG = −64.51 kJ (3)

Mg(OH)2 + 2HF → 2H2 O+ MgF2 ↓, ΔG = −232.14 kJ (4)

The EDS spectrums of the ARRm-F24 h and ARRm-H380-F24 h specimen confirm the presence of
the MgF2 on the surface, moreover, the fluorine amount of ARRm-H380-F24 h is significantly higher

79
Metals 2017, 7, 481

than ARRm-F24 h. ARRm-F24 h possessed only a 1.6 μm fluoride layer, while ARRm-H380-F24 h
possessed a 2.3 μm fluoride layer, as shown in Figure 6e,f, respectively. According to our previous
work, the fluoride coating of the Mg-Zn-Zr series alloy is composed of nano-MgF2 and MgZn2 , with the
distribution and homogeneity of the latter playing a key role in the coating formation mechanism [20].
Interestingly, the heat-treated specimens offered a 1.44-fold higher coating conversion efficiency, and
consequently obtained a thicker coating structure than the non-heat-treated specimens. Generally
speaking, a thick and dense coating offers better anti-corrosion performance.

Figure 6. Surface morphologies, elemental analysis and cross-section profiles of the fluoride coated
surfaces: (a) ARRm-F24 h surface; (b) ARRm-H380-F24 h surface; (c) EDS spectrum acquired from
ARRm-F24 h surface; (d) EDS spectrum acquired from ARRm-H380-F24 h surface; (e) cross-section
profile of ARRm-F24 h; (f) cross-section profile of ARRm-H380-F24 h.

In-vitro electrochemical polarization tests are commonly used to evaluate the corrosion resistance
of biodegradable metals. For the electrochemical reaction of the Mg alloy, the cathodic reaction
mentioned above (Equation (2)) is the water-reduction reaction, which is also closely related to the
driving force of hydrogen gas evolution; meanwhile, the anodic reaction represents the oxidation
driving force of the α-Mg matrix (Equation (1)). Figure 7 shows the typical polarization curves of each
experimental group in the r-SBF solution at 37 ◦ C. The corrosion current density (Icorr ) is the most
important electrochemical parameter, and is often used to calculate polarization resistance (Rp ) [23].
Icorr can be derived via the intersection point of Ecorr and the cathodic extrapolation line. Thereafter,
Rp can be calculated using the Stern-Geary equation, the results of which are listed in Table 3.
Among the ARRm, ARRm-F24 h, ARRm-H380 and ARRm-H380-F24 h groups, the ARRm
unmodified group showed the highest Icorr and lowest Rp of 31.6 μA/cm2 and 1046 Ω·cm2 , respectively,

80
Metals 2017, 7, 481

which indicates that it might encounter severe oxidation and corrosion in a physiological electrolytic
environment. After solid-solution heat treatment, the MgZn2 phases decomposed and dissolved
into the α-Mg matrix of ARRm-H380 (fewer micro-galvanic couples), resulting in a lower Icorr (21.2
μA/cm2 ) than ARRm. Moreover, the significant change in current slope suggests that the ARRm-H380
microstructures had the passivation behavior, which indicates that a protective oxide film formed on
the surface [24]. This protective oxide film can inhibit aggressive ions from penetrating and reacting
with the inner metal surface, thereby reducing the risk of forming hydrogen cavities and releasing
highly-concentrated alkali ions [25]. Both the cathodic and anodic current densities were significantly
reduced in the presence of the MgF2 coating; in particular, ARRm-H380-F24 h offered the highest
corrosion resistance and lowest anodic current density, and featured a wide passivation window (Ebreak
− Ecorr ), indicating that the solid-solution heat treatment could further trigger better coating quality,
performance and the overall anti-corrosion ability.

βa βc
Rp =
2.303(βa + βc ) Icorr
where Icorr is the corrosion current density, while βa and βc are the anodic and cathodic slopes,
respectively, as obtained from the Tafel region.

Figure 7. Electrochemical polarization curves obtained in r-SBF solution at 37 ◦ C.

In our previous work, a duplex modified (microstructure modified and surface coated)
Mg-Zn-Zr alloy was developed [20]. Although this material exhibited improved anti-corrosion
behavior, the importance of the solid-solution heat treatment for fluoride conversion has not been
discussed. To further understand the effect of pre-solid-solution heat treatment, the protection
efficiency percentage (PE%) is employed to elucidate the importance and contribution of this novel
fabrication process.
Icorr,sub − Icorr,MgF2
PE(%) = × 100%
Icorr,sub
The PE value of the fluoride coating grown on the ARRm substrate was found to be smaller
(~80.3%) than that of the ARRm-H380 substrate (~93.8%); this may be due to the latter having superior
coating homogeneity and a thicker coating. It is worth noting that the solid-solution heat treatment
can balance the potential difference between the MgZn2 particles and Mg matrix. Moreover, owing to
hydrogen gas evolution usually occurring at the micro-cathode site, the solid-solution heat treatment
can also prevent the formation of a coating-depletion region on un-dissolved MgZn2 particles, resulting
in higher coating efficiency and protection ability (as shown in Figures 6 and 7).

81
Metals 2017, 7, 481

Table 3. The resultant electrochemical polarization parameters.

Group Ecorr (V) Icorr (μA/cm2 ) βcathodic (V/dec) βanodic (V/dec) Rp (Ω·cm2 ) PE (%)
ARRm −1.44 31.6 0.32 0.10 1046 -
ARRm-F24 h −1.48 6.2 0.25 0.64 12,590 80.3
ARRm-H380 −1.40 21.2 0.23 0.16 1932 -
ARRm-H380-F24 h −1.40 1.3 0.25 0.53 56,739 93.8

3.4. Effect of Clinical Pre-Bending Procedure on Corrosion Rate and Behavior of Mg Regeneration Membrane
Considering the practical application of the dental GBR procedure, dentists need to bend (inducing
strain and residual stress) the regeneration membrane to fit the alveolar ridge [5]. However, Mg-based
alloys generally have a known issue, namely the stress-corrosion cracking (SCC) phenomenon,
which can accelerate localized corrosion and further cause early failure of the materials [26]. Therefore,
the additional effect of bending residual stress on Mg regeneration membranes must be considered
and examined before clinical trial. To realize the effect of bending-corrosion behavior, un-bent/r-SBF
immersion and pre-bent/r-SBF immersion experiments are discussed in the following.
The un-bent/immersion experiment was carried out at 37 ◦ C in r-SBF for 1, 2 and 4 weeks.
Figure 8a shows the corrosion rate data calculated from the normal/immersion test. During the
testing period, the corrosion rates of the modified samples were significantly higher than those
of the unmodified samples, the corrosion-trend sequence of which from fast to slow corrosion is:
ARRm > ARRm-H380 > ARRm-F24 h > ARRm-H380-F24 h. In addition, the corrosion trends of the
immersion and electrochemical tests were identical. Figure 8b shows typical corroded surfaces of
the different samples tested in r-SBF. The surface corrosion morphologies of the ARRm-H380 and
ARRm-H380-F24 h specimens after 1, 3 and 7 days immersion shows the most homogeneous corrosion
morphology. By contrast, localized corrosion, characterized by severe oxidation from the surface to the
interior of the matrix, can be observed in the ARRm and ARRm-F24 h samples.

Figure 8. (a) Corrosion rates calculated from immersion test; (b) corrosion macrograph (red arrow: the
location of corrosion pits).

82
Metals 2017, 7, 481

Figure 9 presents the corrosion macrograph and micrographs of pre-bent/r-SBF immersion


specimens. There are two regions of note on the pre-bent specimens. Firstly, the top surface
of the strained area is where large tensile stress is located. Only the ARRm specimen displays
poor anti-stress corrosion performance, as compared with the other three specimens. Secondly,
due to the heterogeneous contact interface of the nylon screw fixation site, crevice corrosion might
be triggered. According to related reports, Ghali et al. stated that crevice corrosion could be
initiated due to the different hydrolysis rates between the heterogeneous contact interface of Mg
alloys. Accordingly, the formation of Mg(OH)2 could affect the corrosion driving force between the
Mg regeneration membrane/screw interface in the crevices [27]. As seen in Figure 9a, the screw
fixation interface of the ARRm, ARRm-F24 h, and ARRm-H380 specimens display severe crevice
corrosion and accumulation of corrosion-product morphologies; however, the corrosion damage of
the heat-treated sample (ARRm-H380) was clearly less than that of the ARRm and ARRm-F24 h
samples. Notably, ARRm-H380-F24 h showed a satisfactory crevice corrosion-resistant behavior
without significant corrosion damage, confirming that the solid-solution heat treatment can improve
crevice-corrosion resistance.

Figure 9. The pre-bent/immersion test examined specimens: (a) macrograph; and micrograph of
(b) ARRm; (c) ARRm-F24 h; (d) ARRm-H380; (e) ARRm-H380-F24 h.

After immersion in r-SBF for 1 week, the corrosion trends from the un-bent/immersion and
pre-bent/immersion tests were the same (Figure 10); however, the corrosion rates calculated from the

83
Metals 2017, 7, 481

pre-bent/immersion test were higher than those calculated from un-bent/immersion test. This can be
logically explained by the effects of the bending residual stress and heterogeneous interface (screw
fixation site), which can further accelerate the corrosion reaction.

Figure 10. Corrosion rates calculated from un-bent/immersion (marked as normal) and
pre-bent/immersion test (marked as pre-bending).

Relatively higher degradation rates for the non-coated series groups (namely ARRm and
ARRm-H380) were found, which suffered aggressive corrosion by chlorine and/or other reactive
ions. By comparison, after the H380 treatment, the specimens showed relatively lighter corrosion
than the untreated ARRm. Moreover, the results also indicate that the fluoride coatings formed on
the H380-treated specimens are denser and have fewer defects, thereby providing lower corrosion
rates. As such, it is of great significance to reduce the secondary phase of the Mg alloy substrate
(the contribution of H380), in order to improve the coating integrity and its protection against corrosion.
Song et al. reported similar outcomes, where pre-solid-solution treatment could effectively improve
the coating integrity of a Mg-2Zn-Mn-Ca-Ce alloy [28]. Considering that the performance of the
heat-treated samples was superior, the following animal experiments only used the ARRm-H380 and
ARRm-H380-F24 h alloys to examine the practicability.

3.5. In Vivo Degradation and Bone Healing Situation


During the implantation period, all experimental rats showed good health and wound healing
until the end of this research. There were no severe side effects, obvious weight reduction, allergies,
rejection and postoperative infection in the rats. Figure 11 shows the hydrogen accumulation
phenomenon in the ARRm-H380 group after 4 weeks and 12 weeks of implantation. As can
be seen, the photograph shows that ARRm-H380 produced a subcutaneous hydrogen gas cavity,
indicating that the degradation amount of ARRm-H380-F24 h was significantly lower than ARRm-H380.
The generation of hydrogen gas cavity is inevitable, due to the nature of Mg corrosion. This issue
might cause swollen feeling at implanted area. Fortunately, swollen feeling can be minimized by
subcutaneous puncture procedure (to leak the hydrogen cavity).

84
Metals 2017, 7, 481

Figure 11. Hydrogen gas cavity and allergy observation.

As shown in Figure 12, the difference in degradation between ARRm-H380 and ARRm-H380-F24
h lies in the corrosion behavior. Whereas the ARRm-H380 sample showed a homogenous corrosion
morphology, resulting in an evenly corroded cross-section, the ARRm-H380-F24 h specimen showed a
localized corrosion morphology due to the corrosive factors penetrating into the substrate at the weak
points in the fluoride coating [29]. Therefore, the weak points corroded first, after which they evolved
into localized corrosion pits and/or holes. Furthermore, the thickness and weight retention of the
degraded ARRm-H380-F24 h was significantly thicker and heavier than the degraded ARRm-H380,
which means that the degradation and amount of released Mg ions of ARRm-H380-F24 h were
relatively lower than for ARRm-H380.

Figure 12. (a) 3D images of implanted Mg samples; (b) weight retention calculated from CT data.
(arrow: the location of initial corrosion pit).

Figure 13 presents the 3D reconstruction images derived from the μ-CT analysis. The sham group
showed no significant new bone regeneration; a bone fracture of the 5 mm size-level defect can not

85
Metals 2017, 7, 481

be healed because the rat cranial bone lacks blood supply and muscle tissue. However, a remarkable
new bone-regeneration phenomenon can be observed in the ARRm-H380 group after 4 weeks and
12 weeks implantation. In comparison, the ARRm-H380-F24 h group showed relatively lower new
bone-regeneration capability than with ARRm-H380.

Figure 13. (a) 3D images of CSD sites of different experimental groups; (b) new bone volume fraction
calculated from CT data. (Data presented as mean ± SD, n = 4 and analyzed using a one-way ANOVA,
* p < 0.05).

Interestingly, in comparing Figures 12 and 13, it appears that a higher degradation amount of Mg
substrate leads to superior bone regeneration ability. Hence, our results indicate that Mg ions act as an
effective factor for bone growth. Moreover, with respect to the statistical data (Figure 13b), it seems
clear that the bone-regenerative capability of the Mg-based regeneration membrane accelerated bone
tissue formation. The new bone volume fraction with the ARRm-H380 membrane approached almost
100% after 12 weeks implantation. Previous reports have verified that Mg ions act as a stimulator to
enhance cell proliferation and migration, and that the functional biochemical stimulation of Mg ions can
improve wound healing in vitro and in vivo [7,30]. To the best of our knowledge, this report is the first
to discover that Mg-based materials are capable of being applied in dentistry with excellent outcomes.
This suitability could be attributed to the excellent mechanical structuring function and appropriate
degradation properties of modified ECO505 (especially ARRm-H380 and ARRm-H380-F24 h).
In this study, we demonstrated the first report of the promoting effect for bone healing of the
cranial bone in SD rats using modified ECO505 magnesium alloy to acquire the proper magnesium
releasing concentration for bone tissue regeneration. According to the referenced articles and the
study reported here, we created a representative illustration to demonstrate the promoting mechanism
(see Figure 14). These results verified that the ARRm-H380 possessed a proper Mg-releasing ability to
stimulate and enhance the regeneration of the bone defect areas, which is one of the key success factors
for a GBR material. Although ARRm-H380-F24 h showed relatively lower bone-regeneration ability,
its delayed degradation characterization can be effectively applied in older patients who generally
require more recovery time. Therefore, this research successfully provides two future therapy solutions
for different patients or therapy demands, namely short implantation period with ultra-fast healing
effect, and long implantation period with moderate healing effect.

86
Metals 2017, 7, 481

Figure 14. Schematic representation of Mg-based regeneration membrane degradation and new bone
regeneration characteristics.

4. Conclusions
1. The environmentally friendly ECO505 material was successfully developed as the raw material
of a novel Mg-based regeneration membrane.
2. The H380 solid-solution treatment could lead to the recovery of the matrix, refinement and
reduction of the MgZn2 secondary phases while enhancing elongation. Optimization of
the solid-solution heat treatment requires the precisely controlled condition of 380 ◦ C_2 h
to prevent excessive grain growth in the ARRm material, resulting in ultimate elongation
(20.2%). The mechanical properties of the coarse-grained microstructure (ARRm-380 ◦ C_4 h and
ARRm-380 ◦ C_10 h) are not reliable for practical application.
3. ARRm-H380 can form a better fluoride coating quality on substrates than non-heat-treated
ARRm substrate, thereby showing a higher gain of PE value (93.8%) and corrosion resistance
(56 kΩ·cm2 ).
4. ARRm-H380 and ARRm-H380-F24 h specimens can effectively enhance corrosion resistance
and minimize the effects of stress corrosion and crevice corrosion, and so constitutes promising
candidates for regeneration membrane treatment.
5. ARRm-H380 had a significant positive influence on new bone regeneration, where the CSD could
heal almost 100% after 12 weeks implantation. In addition, although ARRm-H380-F24 h showed
relatively lower bone regeneration ability, it nevertheless exhibited better long-term corrosion
resistance than ARRm-H380.

Acknowledgments: The authors are grateful to the Instrument Center of National Cheng Kung University and
Southern Taiwan Medical Device Industry Cluster, AZ-10-16-35-105, for their financial support of this research.
Author Contributions: Da-Jun Lin conceived and designed the experiments; Da-Jun Lin and Hung-Pang Lee
performed the experiments; Da-Jun Lin and Fei-Yi Hung analyzed the data; Fei-Yi Hung and Ming-Long Yeh
contributed reagents/materials/analysis tools; Da-Jun Lin wrote the paper.
Conflicts of Interest: The authors declare no conflict of interest.

87
Metals 2017, 7, 481

References
1. Bottino, M.C.; Thomas, V.; Schmidt, G.; Vohra, Y.K.; Chu, T.M.G.; Kowolik, M.J.; Janowski, G.M.
Recent advances in the development of GTR/GBR membranes for periodontal regeneration—A materials
perspective. Dent. Mater. 2012, 28, 703–721. [CrossRef] [PubMed]
2. Mealey, B.L. Periodontal disease and diabetes—A two-way street. J. Am. Dent. Assoc. 2006, 137, S26–S31.
[CrossRef]
3. Page, R.C.; Altman, L.C.; Ebersole, J.L.; Vandesteen, G.E.; Dahlberg, W.H.; Williams, B.L.; Osterberg, S.K.
Rapidly progressive periodontitis—A distinct clinical condition. J. Periodontol. 1983, 54, 197–209. [CrossRef]
[PubMed]
4. Yoshikawa, G.; Murashima, Y.; Wadachi, R.; Sawada, N.; Suda, H. Guided bone regeneration (GBR) using
membranes and calcium sulphate after apicectomy: A comparative histomorphometrical study. Int. Endod. J.
2002, 35, 255–263. [CrossRef] [PubMed]
5. Rakhmatia, Y.D.; Ayukawa, Y.; Furuhashi, A.; Koyano, K. Current barrier membranes: Titanium mesh and
other membranes for guided bone regeneration in dental applications. J. Prosthodont. Res. 2013, 57, 3–14.
[CrossRef] [PubMed]
6. Witte, F.; Hort, N.; Vogt, C.; Cohen, S.; Kainer, K.U.; Willumeit, R.; Feyerabend, F. Degradable biomaterials
based on magnesium corrosion. Curr. Opin. Solid State Mater. Sci. 2008, 12, 63–72. [CrossRef]
7. Lin, D.J.; Hung, F.Y.; Yeh, M.L.; Lui, T.S. Microstructure-modified biodegradable magnesium alloy for
promoting cytocompatibility and wound healing in vitro. J. Mater. Sci. Mater. Med. 2015, 26, 1–10. [CrossRef]
[PubMed]
8. Lin, D.J.; Hung, F.Y.; Yeh, M.L.; Lee, H.P.; Lui, T.S. Development of a novel micro-textured surface using
duplex surface modification for biomedical Mg alloy applications. Mater. Lett. 2017, 206, 9–12. [CrossRef]
9. Xin, Y.; Hu, T.; Chu, P.K. In vitro studies of biomedical magnesium alloys in a simulated physiological
environment: A review. Acta Biomater. 2011, 7, 1452–1459. [CrossRef] [PubMed]
10. Lin, D.J.; Hung, F.Y.; Jakfar, S.; Yeh, M.L. Tailored coating chemistry and interfacial properties for construction
of bioactive ceramic coatings on magnesium biomaterial. Mater. Des. 2016, 89, 235–244. [CrossRef]
11. Cai, S.H.; Lei, T.; Li, N.F.; Feng, F.F. Effects of Zn on microstructure, mechanical properties and corrosion
behavior of Mg-Zn alloys. Mater. Sci. Eng. C 2012, 32, 2570–2577. [CrossRef]
12. Song, Y.W.; Han, E.H.; Shan, D.Y.; Yim, C.D.; You, B.S. The role of second phases in the corrosion behavior of
Mg-5Zn alloy. Corros. Sci. 2012, 60, 238–245. [CrossRef]
13. Lin, D.J.; Hung, F.Y.; Lui, T.S.; Yeh, M.L. Heat treatment mechanism and biodegradable characteristics of
ZAX1330 mg alloy. Mater. Sci. Eng. C 2015, 51, 300–308. [CrossRef] [PubMed]
14. Yan, T.T.; Tan, L.L.; Xiong, D.S.; Liu, X.J.; Zhang, B.C.; Yang, K. Fluoride treatment and in vitro corrosion
behavior of an AZ31B magnesium alloy. Mater. Sci. Eng. C 2010, 30, 740–748. [CrossRef]
15. Bakhsheshi-Rad, H.R.; Idris, M.H.; Kadir, M.R.A.; Daroonparvar, M. Effect of fluoride treatment on corrosion
behavior of Mg-Ca binary alloy for implant application. Trans. Nonferr. Met. Soc. China 2013, 23, 699–710.
[CrossRef]
16. Lin, D.J.; Hung, F.Y.; Liu, H.J.; Yeh, M.L. Dynamic corrosion and material characteristics of Mg-Zn-Zr
mini-tubes: The influence of microstructures and extrusion parameters. Adv. Eng. Mater. 2017. [CrossRef]
17. Li, Z.; Shizhao, S.; Chen, M.; Fahlman, B.D.; Debao, L.; Bi, H. In vitro and in vivo corrosion, mechanical
properties and biocompatibility evaluation of MgF2-coated Mg-Zn-Zr alloy as cancellous screws. Mater. Sci.
Eng. C 2017, 75, 1268–1280. [CrossRef] [PubMed]
18. Ramakrishnan, S.; Koltun, P. Global warming impact of the magnesium produced in china using the pidgeon
process. Resour. Conserv. Recycl. 2004, 42, 49–64. [CrossRef]
19. Ha, W.; Kim, Y.J. Effects of cover gases on melt protection of Mg alloys. J. Alloy. Compd. 2006, 422, 208–213.
[CrossRef]
20. Lin, D.J.; Hung, F.Y.; Yeh, M.L.; Lee, H.P.; Lui, T.S. Correlation between anti-corrosion performance and
optical reflectance of nano fluoride film on biodegradable Mg-Zn-Zr alloy: A non-destructive evaluation
approach. Int. J. Electrochem. Sci. 2017, 12, 3614–3634. [CrossRef]
21. Prasad, Y.V.R.K.; Rao, K.P. Effect of homogenization on the hot deformation behavior of cast AZ31 magnesium
alloy. Mater. Des. 2009, 30, 3723–3730. [CrossRef]

88
Metals 2017, 7, 481

22. Watanabe, H.; Mukai, T.; Ishikawa, K.; Higashi, K. Low temperature superplasticity of a fine-grained ZK60
magnesium alloy processed by equal-channel-angular extrusion. Scr. Mater. 2002, 46, 851–856. [CrossRef]
23. Stern, M.; Geary, A.L. Electrochemical polarization-1. A theoretical analysis of the shape of polarization
curves. J. Electrochem. Soc. 1956, 103, C205.
24. Chang, J.W.; Guo, X.W.; Fu, P.H.; Peng, L.M.; Ding, W.J. Effect of heat treatment on corrosion and
electrochemical behaviour of Mg-3Nd-0.2Zn-0.4Zr (wt. %) alloy. Electrochim. Acta 2007, 52, 3160–3167.
[CrossRef]
25. Lee, H.P.; Lin, D.J.; Yeh, M.L. Phenolic modified ceramic coating on biodegradable Mg alloy: The improved
corrosion resistance and osteoblast-like cell activity. Materials 2017, 10. [CrossRef] [PubMed]
26. Winzer, N.; Atrens, A.; Song, G.L.; Ghali, E.; Dietzel, W.; Kainer, K.U.; Hort, N.; Blawert, C. A critical review
of the stress corrosion cracking (SCC) of magnesium alloys. Adv. Eng. Mater. 2005, 7, 659–693. [CrossRef]
27. Ghali, E.; Dietzel, W.; Kainer, K.U. Testing of general and localized corrosion of magnesium alloys: A critical
review. J. Mater. Eng. Perform. 2004, 13, 517–529. [CrossRef]
28. Song, D.; Li, C.; Zhang, L.; Ma, X.; Guo, G.; Zhang, F.; Jiang, J.; Ma, A. Decreasing bio-degradation rate of the
hydrothermal-synthesizing coated Mg alloy via pre-solid-solution treatment. Materials 2017, 10. [CrossRef]
[PubMed]
29. Witte, F.; Fischer, J.; Nellesen, J.; Vogt, C.; Vogt, J.; Donath, T.; Beckmann, F. In vivo corrosion and corrosion
protection of magnesium alloy LAE442. Acta Biomater. 2010, 6, 1792–1799. [CrossRef] [PubMed]
30. Schaller, B.; Saulacic, N.; Imwinkelried, T.; Beck, S.; Liu, E.W.Y.; Gralla, J.; Nakahara, K.; Hofstetter, W.;
Iizuka, T. In vivo degradation of magnesium plate/screw osteosynthesis implant systems: Soft and hard
tissue response in a calvarial model in miniature pigs. J. Cranio-Maxillofac. Surg. 2016, 44, 309–317. [CrossRef]
[PubMed]

© 2017 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

89
metals
Article
Investigation on Mechanical Behavior of
Biodegradable Iron Foams under Different
Compression Test Conditions
Reza Alavi, Adhitya Trenggono, Sébastien Champagne and Hendra Hermawan *
Department of Mining, Metallurgical and Materials Engineering and Centre Hospitalier Universitaire de Québec
Research Center, Laval University, Quebec City, QC G1V 0A6, Canada; reza.alavi.1@ulaval.ca (R.A.);
adhitya.trenggono.1@ulaval.ca (A.T.); sebastien.champagne.2@ulaval.ca (S.C.)
* Correspondence: hendra.hermawan@gmn.ulaval.ca; Tel.: +1-418-656-2131 (ext. 5876)

Academic Editor: Eli Aghion


Received: 13 April 2017; Accepted: 30 May 2017; Published: 2 June 2017

Abstract: Biodegradable metal foams have been studied as potential materials for bone scaffolds.
Their mechanical properties largely depend on the relative density and micro-structural geometry.
In this work, mechanical behavior of iron foams with different cell sizes was investigated under
various compression tests in dry and wet conditions and after subjected to degradation in
Hanks’ solution. Statistical analysis was performed using hypothesis and non-parametric tests.
The deformation behavior of the foams under compression was also evaluated. Results show that
the mechanical properties of the foams under dry compression tests had a “V-type” variation,
which is explained as a function of different geometrical properties by using a simple tabular method.
The wet environment did not change the compression behavior of the iron foams significantly
while degradation decreased the elastic modulus, yield and compression strengths and the energy
absorbability of the specimens. The deformation of open cell iron foams under compression is viewed
as a complex phenomenon which could be the product of multiple mechanism such as bending,
buckling and torsion.

Keywords: biodegradable metals; iron foam; scaffold; compression; degradation; cell size; open
cell foam

1. Introduction
Porous metals or metal foams are used in different applications where altered material properties
of the parent metal are beneficial to the quality of the application. In biomedical engineering, metal
foams can be used for biodegradable orthopedic implants such as bone scaffolds. Although foaming
does not change all the material properties of the cell-wall material, there are properties which
depend on the density (and therefore porosity) and micro-structural geometry of cellular materials:
“the stiffness, the mechanical strength, the thermal and electrical conductivity as well as acoustic
properties” [1]. One should note that the stiffness (elastic modulus) of a cellular structure, e.g., iron
foam, depends on the architecture of the structure to a great extent. Therefore, it should not be confused
with the elastic modulus of the cell wall material, e.g., iron [1]. As stated in [2], the material properties
of metal foams “most directly” depend on the relative density and the properties of the parent material.
However, structural properties such as pore sizes, cell types, etc. also influence the material properties
of the foams. Bone scaffolds are implanted in the body to serve as a platform on which bone formation
takes place. An ideal bone scaffold would resemble the mechanical properties of the natural bone.
One advantage of porous metals over solid metals is their lower stiffness, which makes it closer to that
of a bone. This would help to reduce stress shielding which can happen in case of using solid metals

Metals 2017, 7, 202; doi:10.3390/met7060202 90 www.mdpi.com/journal/metals


Metals 2017, 7, 202

due to their higher Young’s modulus than that of the bone. Porous structure can also contribute to new
tissue infiltration and bone formation [3,4]. The structure of the porous scaffold (pore size distribution,
interconnectivity and porosity) is an influential factor on the quality of the scaffold as it affects the
level of cell penetration, cell growth and material transportation into and out of the scaffold [5].
Different biocompatible porous metals have been studied to serve as scaffolds for orthopedic
applications: tantalum, magnesium and its alloys, titanium and its alloys, and iron-foam based
materials [5,6]. Magnesium, iron and zinc are considered as biodegradable metals. However, among the
available studies on biodegradable systems, the majority of the investigations have been on magnesium
based materials due to their non-toxicity and similar mechanical behavior to that of human bone [7].
There have been controversies over biocompatibility of iron due to the emergence of metallosis, local
destruction of tissues as a result of mechanical-biological, electro-energetic, and chemical-toxic effects
of metal after implantation, of iron implants [8]. Nevertheless, there exist studies that show iron-based
materials are possibly suitable for temporary biodegradable implants [9,10]. They provide answers to
the two major drawbacks of magnesium-based materials which are high degradation rate that limits
the use of such materials on small implants with approximate life span of 6–12 months, and hydrogen
evolution during corrosion that can disturb the healing process [7,11–13]. Mechanical properties
of iron-based alloys, e.g., strength and ductility, can be easily tailored to meet the criteria for some
biomedical applications. They are viewed as good candidates for load bearing biodegradable implants
owing to their high mechanical properties, e.g., high strength [13], and biocompatible, non-toxic
characteristics [7]. However, a major challenge of using iron as a biodegradable implant is its slow rate
of degradation [7,14]. Different approaches have been proposed to improve the corrosion rate of iron
such as alloying elements modification, poly(lactic-co-glycolic acid) infiltration, and coating [6,15–17].
The objective of this study is to investigate the mechanical behavior of iron-foams as a function of their
structural geometry where several iron foam specimens with different structural properties underwent
uniaxial compression tests. Results are discussed in terms of the influence of porous properties, i.e., cell
size, pore size, number of pores and strut thickness, and influence of environmental conditions, i.e., wet
and dry condition as well as degradation on mechanical behavior of iron foam samples. Deformation
mechanisms of iron foams under compression were also studied using scanning electron microscope
images (SEM) analysis.

2. Materials and Methods

2.1. Iron Foam Specimens


The specimens were open cell pure iron foam with nominal cell diameters of 450 (IF45), 580 (IF58)
and 800 (IF80) μm manufactured by Alantum. The foams were produced by the replication of open
cell polyurethane (PU) foams [18]. To do so, a thin layer of Ni is sputtered on the PU foam to make
the foam conductive for the following electroplating process. Then, iron would be electroplated on
the foam to produce the open cell iron foams [18]. Measurements of cell sizes, pore sizes and strut
thicknesses of iron foam samples were performed using scanning electron microscope images (SEM,
Quanta 250 FEI, Hillsboro, OR, USA). More details on the geometrical measurement are available
in Section A of the Supplementary Materials (Figure S1 and Table S1). One should note that these
measurements were conducted on 2D images, and the 3D structure of the foams was not considered.
Thus, the average measurements should be considered as estimations. The relative density of a foam is
defined as ratio of the foam density over the density of the cell-wall material (ρ*/ρs ) [19]. To obtain the
foam densities, the iron foam sheets were cut into cubic specimens by a stainless-steel scalpel. For each
group of iron foams, three specimens were used. The dimensions of the specimens were measured via
caliper to obtain the apparent volume (the bulk volume which contains the struts and pores). The mass
of each specimen was measured via a sensitive digital scale. The foam densities were calculated as
the ratio of masses over apparent volumes. The density of the cell-wall material (iron) was taken as
7.874 (g/cm3 ). The calculated average relative density values of IF45, IF58 and IF80 samples were

91
Metals 2017, 7, 202

0.038, 0.027 and 0.025, respectively (Table 1). It should be mentioned that for calculation of the relative
densities, it was assumed that the solid structures of the foams contained only iron, and no remainder
of PU was present. To prepare the cubic samples for the mechanical tests, they were initially cut from
the sheets via a stainless-steel scalpel slightly larger than the final dimensions, and the thicknesses
remained unchanged. Then the width and length were reduced to the final dimension by a rotary
cutting tool kit. The nominal width and length of iron foam specimens considered for dry compression
tests was 10 mm × 10 mm, and the nominal thickness (along the loading direction) for IF45, IF58 and
IF80 was 1.7, 2.1 and 2.6 mm, respectively. According to [20], the minimum requirement for the size of
all the spatial dimensions of the specimens and for their ratio to the average pore size is 10 mm and 10,
respectively. The length and width of the iron foam specimens satisfy both mentioned requirements
and have been verified as detailed in Section B of the Supplementary Materials (Figure S2). For wet
compression tests and immersion tests, the specimens (only IF80) had similar nominal dimensions as
the dry test specimens. To understand the deformation mechanisms, a new set of IF45 specimens were
prepared to undergo dry compression tests before SEM observations. They were also prepared in the
same fashion as other specimens with similar nominal dimensions. However, because no quantitative
analysis was involved in this part, the accuracy of measured dimensions was not critical.

2.2. Mechanical Testing

2.2.1. Compression Test Parameters and Conditions


Figure 1 presents typical stress–strain curves of metallic foams under compression. Different
specimen sizes and cross-head speeds were considered initially to choose proper test parameters.
Finally, it was decided to use specimens with compression area of 100 mm2 (A = 10 × 10 mm2 ) and
the cross-head speed 0.001 mm/s for all the tests and analyses. The compression tests under dry and
wet conditions were carried out by Instron machine (ElectroPuls E1000, Instron, Norwood, MA, USA)
with a 2 kN load cell. The details of the procedure for choosing the test parameters as well as the
dry compression stress–strain curves of all the iron foam samples are available in Section B of the
Supplementary Materials (Figures S2–S5).

Figure 1. (a) A typical stress–strain curve of a foam under compression, numbers indicate different
methods to define the compression strength; and (b) determination of densification strain. Adapted
from [21,22].

Given that the biological environment within the body is not dry, wet compression tests were
carried out on IF80 specimens to provide a more realistic condition. The tests were conducted within
Hanks’ solution (H1387, Sigma Aldrich, Saint Louis, MO, USA) at 37 ◦ C filled into a cylindrical
bath (inside diameter = 140 mm, solution fill height = 50 mm) mounted on the Instron machine
after the specimen is fixed without gap within the platens (diameter = 50 mm, thickness = 25 mm)

92
Metals 2017, 7, 202

and before loading. The configuration allowed the solution to escape during compression and
hydrostatic pressure was avoided. In addition, considering that the environment within the body
is corrosive, static immersion tests were conducted on IF80 specimens to investigate the effects of
degradation on mechanical behavior of iron foams. In the static immersion, the specimens were hung
from a wire and submerged in 100 mL of the Hanks’ solution then placed in separate incubators
(ThermoFisher Scientific, Waltham, MI, USA) for three or seven days at 37 ◦ C and 5% CO2 atmosphere.
The compression test parameters and the specimen nominal sizes were identical to those of the
previous dry and wet tests. Finally, a few compression tests were performed on IF45 specimens in
order to understand the deformation and failure mechanism of iron foams under compression. To do
so, the specimens underwent compression up to different strain levels and then they were observed
under SEM.

2.2.2. Compression Properties


In order to assess the mechanical behavior of the iron foams, the following properties were
determined using the stress–strain response of the specimens: Elastic modulus (E), yield strength
(σy ), compression strength (σc ), densification strain (εD ) and the energy absorbability per volume
up to the point of densification (W). The elastic modulus values were approximated by the linear
fitting tool of the Quick Fit Gadget provided in OriginPro 2016 software (OriginLab, Northampton,
MA, USA). After estimating the slope of the linear elastic regime (E) on each stress–strain curve,
the yield strength was approximated by 0.2% offset method [23]. For most of the iron foam specimens
investigated in this study, compression strength (σc ) was taken as the first local maximum after elastic
regime (marked by 2 and σc in Figure 1a,b, respectively). However, if there was no apparent first local
maximum, an arbitrary local maximum in the plateau region was chosen to represent the compression
strength stated by Banhart and Baumeister [21]. Densification strain (εD ) can be defined as a strain
at which densification begins. However, because normally there is no abrupt transition from plateau
regime to densification regime, εD can be defined as the intersection of the tangents to the plateau
and densification regimes (marked as εD in Figure 1b) [22]. In this study, it was tried to draw the
tangents from the midpoint of the plateau regime (where the second derivation is close to zero) and
from a part of the densification regime where the slope seems to become stable. The points were
selected by eyeballing. A larger εD implies that the material undergoes higher strains before the onset
of densification. The value of energy absorbed per volume up to the densification strain (W) is equal to
the area underneath the stress–strain diagram from ε = 0 to ε = εD , expressed as [22]:
 εD
W= σ(ε)dε (1)
0

The point of ε = 0 was determined as the intersection of the line continuing the linear elastic
regime with the strain axis. This method was inspired from the definition of the “zero point for the
compressive strain” provided in [20]. High energy absorption capacity indicates the higher level
of impact absorption by the material. However, in order to study the impact behavior in a more
comprehensive fashion, conducting dynamic impact tests with higher strain rates beside static tests is
recommended [24].

2.3. Statistical Analysis


Values of E, σy, σc , εD and W were compared against one another for different groups of specimens
via statistical analysis and were shown as mean ± standard deviation. In order to draw a reliable
conclusion, hypothesis tests (One-way ANOVA, t-test) along with non-parametric tests (Kruskal–Wallis
ANOVA, Two-Sample Kolmogorov–Smirnov Test) were carried out in the OriginLab software.
The non-parametric tests were conducted due to the small sample sizes. Unlike parametric hypothesis
tests, the normal distribution of population is not assumed when conducting non-parametric tests [25].

93
Metals 2017, 7, 202

3. Results

3.1. Iron Foam Structure


Figure 2 depicts the structure of iron foams with different average cell sizes. It shows that IF80
specimens tend to have thicker struts than those of IF45 and IF58 specimens. Strut thicknesses of
the iron foams of different cell sizes are shown in Table 1. The mean values of both branch-strut and
end-strut thicknesses of the IF80 specimen are significantly higher than those of the IF45 and IF58
specimens. Definition of cell size, branch-strut and end-strut as well as detail measurement are given
in Figure S1 and Table S1 of the Supplementary Materials.

Figure 2. Scanning electron microscope (SEM) images of iron foam structures: (a) IF45; (b) IF58; and
(c) IF80.

Table 1. Relative density, cell size, end-strut and branch-strut thicknesses of the iron foams.

Relative Branch-Strut
Specimen Cell Size (μm) Pore Size (μm) * End-Strut (μm)
Density (μm)
IF45 0.038 ± 0.001 461.77 ± 72.26 155.59± 27.94 74.73 ± 10.30 55.52 ± 6.18
IF58 0.027 ± 0.001 617.73 ± 76.08 150.8 ± 29.43 63.62 ± 9.95 59.88 ± 7.55
IF80 0.025 ± 0.001 828.11 ± 79.87 157.33 ± 28.50 97.79 ± 17.54 80.91 ± 12.27
* Between 100 and 200 μm.

3.2. Dry Compression Behavior of the Iron Foams


The stress–strain curves resulted from the dry compression tests on iron foams of different cell
sizes are shown in Figure 3. The shifting of the curves within a sample group can be the result of
non-identical micro configurations even if they are in the same sample group. The IF45 specimens
tend to have higher compression strength than those of IF58 and IF80 specimens. Strain hardening
up to the peak followed by a softening is more visible in the stress–strain curves of IF45 and IF80
specimens than those of IF58 specimens. Thus, the local maximum after the linear elastic regime
followed by a local minimum in IF58 curves does not stand out as much as it does in IF45 and IF80
curves. The IF58 specimens seem to experience a shorter plateau region than that of IF45 and IF80
specimens. In general, IF45 specimens tend to have the highest strength under compression as they
experience the highest level of stress in all regimes. The compression properties of the iron foams are
summarized in Table 2. There exists a “V-type” variation of the mechanical properties with respect to
the cell sizes, i.e., the mean values of the compression properties of the iron foams of 580 μm nominal
cell sizes tend to be lower than those of the specimens with 450 and 800 μm nominal cell size.

94
Metals 2017, 7, 202

Figure 3. The stress–strain curves of the dry compression tests for different cell sizes: (a) IF45; (b) IF58;
and (c) IF80.

Table 2. Compression properties of iron foam specimens tested in dry condition.

Specimen E (MPa) σy (MPa) σc (MPa) εD (mm/mm) W (MJ/m3 )


IF45 11.60 ± 1.39 0.48 ± 0.07 0.53 ± 0.05 15.03 ± 0.91 0.054 ± 0.009
IF58 8.24 ± 0.67 0.23 ± 0.03 0.26 ± 0.02 9.94 ± 0.54 0.016 ± 0.003
IF80 17.11 ± 2.3 0.36 ± 0.03 0.41 ± 0.04 12.52 ± 0.48 0.033 ± 0.004

3.3. Compression Behavior of Iron Foams in Hanks’ Solution and after Degradation
Elastic modulus, yield and compression strength, densification strain and energy absorbability
of five IF80 specimens under wet compression tests were calculated and compared against those of
IF80 specimens which had undergone dry compression. The properties of specimens after three- and
seven-day immersions were also compared with those which had not undergone any immersion.
The stress–strain diagrams and mean values of compression properties are demonstrated in Figure 4
and Table 3, respectively. None of the properties are significantly affected by the wet environment.
Differently, the compression behavior of IF80 iron foams was affected after three- and seven-day
immersions in Hanks’ solution as shown in Figure 4c,d and in Table 3. All the calculated compression
strength decreased as a result of degradation. However, these reductions are not significant for
densification strain and between the three and seven days after immersion samples.

Table 3. Compression properties of IF80 specimens under dry and wet condition, and after
immersion tests.

Specimen E (MPa) σy (MPa) σc (MPa) εD (%) W (MJ/m3 )


Wet condition 14.14 ± 1.39 0.33 ± 0.44 0.37 ± 0.04 12.49 ± 0.27 0.030 ± 0.003
No immersion 14.78 ± 2.28 0.39 ± 0.04 0.43 ± 0.04 13.21 ± 0.96 0.039 ± 0.004
3-day immersion 10.48 ± 1.39 0.25 ± 0.03 0.31 ± 0.04 13.17 ± 0.50 0.027 ± 0.005
7-day immersion 10.06 ± 1.49 0.25 ± 0.03 0.30 ± 0.03 13.09 ± 0.73 0.025 ± 0.004

95
Metals 2017, 7, 202

Figure 4. The stress–strain curves of IF80 specimens: (a) no immersion; (b) wet conditions; (c) three
days after immersion; and (d) seven days after immersion.

3.4. Statistical Analysis


Both hypothesis tests (ANOVA and Tukey) and non-parametric tests were employed to investigate
the significance of differences among the values of the elastic and plastic properties of the different
iron foams. Figure 5a,b represents the elastic modulus (E) and yield strength (σy ) box-charts of the
specimens under compression, respectively. Although both of the properties represent a V-type
variation, an interesting difference is observed between the two: highest elastic modulus is presented
by IF80 specimens, while the highest yield strength is presented by IF45 specimens. Indeed, among
all the determined properties, it is only the elastic modulus that has a V-type variation in which IF80
specimens present the highest values. Figure 5c depicts the compression strength box-charts of the iron
foam specimens of different cell sizes. The figure shows a V-type variation of the strength (Figure 5c).
The mean values of compression strengths within IF45, IF58 and IF80 samples were 0.53, 0.255 and
0.41 MPa, respectively. According to the ANOVA and Tukey tests, at α = 0.05, all the population means
of σy , E and σc were significantly different from one another. The results obtained from non-parametric
tests agreed with those of hypothesis tests: the null hypothesis in Kruskal–Wallis ANOVA test was
rejected at significance level of 0.05. Two-sample Kolmogorov–Smirnov (K–S) tests with significance
level of 0.05 were carried out within each possible sample pair. The results of K–S tests showed that all
the distributions were significantly different from one another, except for the case of the yield strength
difference between IF45 and IF80 samples which contradicts the result drawn from the Tukey test.
However, given the difference shown by the box-chart of Figure 5b, the Tukey test result seems to
be more reliable, i.e., the difference between the yield strength value of IF45 and IF80 populations
is significant.

96
Metals 2017, 7, 202

Figure 5. Box charts for iron foam specimens of different cell sizes: (a) elastic modulus; (b) yield
strength; (c) compression strength; (d) densification strain and (e) energy absorbability.

Figure 5d depicts the densification strain (εD ) box-charts of the iron foam specimens of different
cell sizes. Similar to the compression strength variations, IF45 and IF58 samples show the highest and
the lowest values of densification strain, respectively (V-type variation). Both One-way ANOVA and
non-parametric Kruskal–Wallis ANOVA tests were conducted to evaluate the significance of difference
between the populations. Based on the ANOVA and Tukey test results, all the population means were
significantly different form one another with a significance level of 0.05. This was in agreement with
the result obtained by Kruskal–Wallis ANOVA test at α = 0.05, and the three two-sample K–S tests
between pairs at α = 0.05. Therefore, it can be concluded that IF45 and IF58 samples possess the highest
and the lowest level of densification strain, respectively. Figure 5e depicts the amount of absorbed
energy per Volume (W) up to the densification strain for specimens of different cell sizes which
was also compared against one another via one-way ANOVA, Tukey, Kruskal–Wallis ANOVA, and
Two-Sample Kolmogorov–Smirnov (K–S) tests. The mean value of energy absorbability within IF45,
IF58 and IF80 samples was 0.054, 0.016 and 0.033 MJ/m3 , respectively. According to the Homogeneity
of Variance test results, the population variances were significantly different (α = 0.05) which violates
the equality-of-variance requirement to perform ANOVA test. Although the ANOVA test results stated
that at least two populations are significantly different, three two-sample t-tests at significance level of
0.01 were performed due to the violation. The results of the t-tests suggest that every population is
significantly different from the other two (α = 0.01). Same conclusion was drawn from performing a
Kruskal–Wallis ANOVA test in conjunction with three two-sample Kolmogorov–Smirnov (K–S) tests
at significance level of 0.05. Therefore, it is concluded that iron foams of 450 μm and 580 μm average
cell sizes have the highest and the lowest level of energy absorbability, respectively.
To compare the results of wet and dry compression tests, both two-sample t-test and two-sample
non-parametric Kolmogorov–Smirnov test at significance level of 0.05 were utilized. For the case of
the elastic modulus, E, the t-test showed that the difference between the values of the two groups
was significant which disagreed with what the Kolmogorov–Smirnov test suggested, i.e., the two
distributions were not significantly different. Thus, considering that the values of E were just an
estimation and that the sample sizes were small, it is not feasible to make a reliable comment on the
elastic behavior of the iron foams within the wet environment. However, for the other compression
properties the results of the statistical tests were consistent. Although the results of compression tests

97
Metals 2017, 7, 202

showed that under wet condition, the mean values of the σy , σc , εD and W tended to be slightly lower
than those of the dry tests, the results of all the hypothesis and non-parametric tests showed otherwise:
None of the calculated mechanical properties under wet condition were significantly different from
those of dry compression. This suggests that the wet environment did not significantly influence the
mentioned mechanical properties of the iron foams.
The significance of difference between the mechanical properties of the iron foams after different
periods of immersion was investigated via hypothesis and non-parametric tests. Figure 6a–c,e depicts
the elastic modulus, yield strength, compression strength and the energy absorbability box charts of
IF80 specimens after three- and seven-day immersion tests, respectively. In addition, the box charts
samples upon which no immersion test was performed are included. The figure shows that there is a
relatively high reduction in the levels of all the four properties after three-day immersion (Figure 6).
In addition, the result of one-way ANOVA indicated that (α = 0.05), at least two sample population
were significantly different, and the results of Tukey test showed that (α = 0.05) the mean differences
between “three-day immersion” and “seven-day immersion” samples were not significant unlike the
other two comparisons, i.e., “no immersion–three-day immersion” and ”no immersion–seven-day
immersion” samples. Identical conclusions were drawn from conducting Kruskal–Wallis ANOVA tests
in conjunction with three Kolmogorov–Smirnov tests within the sample pairs at α = 0.05.

Figure 6. Box charts for iron foam specimens of different immersion time: (a) elastic modulus; (b) yield
strength; (c) compression strength; (d) densification strain and (e) energy absorbability.

Figure 6d depicts the densification strain box-charts of IF80 iron foam specimens after three- and
seven-day immersions as well as those on which no immersion test was performed. Looking at the
box-charts, the difference between the densification strain levels does not seem to be significant. As the
normality of the densification strain values within each sample group was evaluated by Shapiro–Wilk
tests, it appeared that the normality of “seven-day immersion” sample population was rejected unlike
the other two groups. The result of one-way ANOVA and Kruskal–Wallis ANOVA showed that at
α = 0.05, the difference between the sample population were not significant. Although the normality
requirement for the ANOVA test was violated, the result of the test is expected to be valid given the
proximity of strain densification mean values of the samples.

98
Metals 2017, 7, 202

3.5. Deformation of the Iron Foams after Compression


To understand the deformation and failure mechanism of open cell iron foams during compression,
various compression tests were conducted on IF45 specimens followed by SEM observation of the
compressed specimens. Figure 7 depicts four specimens compressed up to strain levels of 10.8%, 12.8%,
29.8% and 49.5%. The areas marked with red circles depict the struts that experienced bending or
plastic deformation. Because the direction of the compression force is perpendicular to the image
surfaces, it is difficult to identify buckling in the struts, if it existed. However, considering that the
majority of the struts are not completely perpendicular or parallel to the loading direction, it is expected
that both bending and buckling contribute to the deformation of some struts.

Figure 7. Deformed structure of four IF45 specimens at compression strain of: (a) 10.8%; (b) 12.8%;
(c) 29.8%; and (d) 49.5%.

Figure 8 depicts the magnified images of some marked regions in Figure 7. It shows the
deformation of struts of four specimens at different stages of compression after elastic regime.
In samples with the maximum compression strain levels of 10.8%, 12.8% and 29.8%, formation of
S-shape plastic hinges in some of the marked regions can be observed, e.g., region 2, 3, 7, 8, and 14
(Figure 8a). C-shape bending is observed in some of the struts of specimens with compression strains
of 29.8% and 49.5%, e.g., region 13, 24 and 26 (Figure 8b). Plastic S-shape and C-shape deformation
in some struts of open cell aluminum alloy (A356) and 316L stainless steel foams under quasi-static
compression has been observed in the previous works by others [26,27]. Deformation bands are present
in a few struts of the specimen with the highest level of compression strain (49.5%). These adjacent
struts are marked with red circles of number 21, 23 and 27 (Figure 8c).

99
Metals 2017, 7, 202

Figure 8. Different shapes formed on plastically deformed struts: (a) S-shape; (b) C-shape;
and (c) deformation bands.

3.6. Morphology of the Iron Foams after Immersion Tests


Figure 9 depicts the morphological structure of three specimens, one of which was not immersed
and the other two were immersed in Hanks’ solution for three and seven days. As shown in the picture,
layers of corrosion products were formed on the struts after immersion. The figure suggests that the
layers on seven-day immersed specimen are thicker than those on three-day immersed specimens.

Figure 9. Morphology of the IF80 samples: (a) no immersion test; (b) after three-day immersion; and
(c) after seven-day immersion.

100
Metals 2017, 7, 202

4. Discussion

4.1. Effect of Structural Properties on Elastic and Plastic Compression Behavior of Iron Foams
As mentioned in the previous sections, stress–strain curves of iron foams under compression
are composed of three main regimes: linear elastic, plateau and densification. In the case of metal
foams, or other foams made of materials with a plastic yield point, as the stress goes beyond the linear
elastic region, plastic collapse takes place. The collapse of cell walls continues until the cells collapse to
an extent which further strain requires much higher level of stress, starting the densification part of
the compression diagrams [28]. However, the presence of some plastic deformations during elastic
regime has been pointed out in compression of some closed-cell aluminum foams [29]. In addition,
a significant softening after hardening was observed in many of the iron foam specimens (Figure 3).
As explained in [29], softening in closed-cell aluminum foams that underwent compressive loading
and unloading was attributed to the collapse of the cells resulted from the plastic collapse of a single
deformation band perpendicular to the loading direction. Similarly, plastic collapse of struts/cells
may have caused the softening in some of the iron foam specimens. Thus, absence of a significant
softening indicates that massive cell collapses do not take place immediately after the beginning of
the plateau. Lower densification strain of IF58 specimens (Table 2) could be the result of smaller
end-strut thicknesses (Table S1) in IF58 specimens, making the cells weaker especially on the loaded
ends, leading them to a faster densification.
In general, the mechanical properties of foams depend on structural properties and the properties
of the cell-wall (parent) material [28]. The most important structural properties are relative density (the
ratio of the density of the cellular material to the density of the solid material), cell type (open or close)
and the level of anisotropy in the cells. The most influential cell wall properties are the solid density (ρs ),
the Young’s modulus (Es ), the yield strength (σys ), the fracture strength (σfs ) and the creep parameters
.
(ns , os and σos ) [28]. Because the cell-wall material properties of the samples are not significantly
different in this study, it is expected that at least some of the geometrical and structural properties such
as the relative densities, cell size, pore size and strut thicknesses influence the mechanical behavior
of the iron foams. In the work of Amsterdam et al. [30], it was observed that the relative density
influenced the plastic collapse stress values of open cell aluminum foams. In this study, however,
the relative density of IF58 and IF80 specimens are very close, so it is expected that other parameters
have had more contribution to the variation of the level of the mechanical properties between the two
samples. The influence of cell size on mechanical behavior of the foams has been a matter of controversy.
As stated in [29], in most cases, mechanical properties of metal foams do not depend on cell size.
This was observed in the study of the deformation behavior of the open-cell stainless steel conducted
by Kaya and Fleck [27], wherein, at the same relative density, the inhomogeneity in the microstructure
was found the influential factor and not the cell size. Investigating the influence of density, cell size
and cell shape on the mechanical properties of open cell 6101 aluminum foams, Nieh et al [31] found
that, under similar densities, cell size does not have any significant effect on strength while the cell
shape had some influences [31]. On the other hand, there have been researchers who found the cell
sizes would affect the mechanical properties of the foam or that of a lattice structure [32,33].
In their work, Jian et al. [34] observed that the compressive and fracture strength of porous NiTi
alloy samples increased with decreasing mean pore size. The samples of different porosities and pore
sizes went under quasi-static compression tests (crosshead speed of 2.4 mm/min) while the range
of porosities and pore sizes were 53–55.6% and around 264.8–1026.6 μm, respectively. Therefore,
the variation between pore size values seems to be much more considerable than that of the porosities.
However, studying the effect of pore size on the mechanical properties of open cell aluminum foams
with spherical pores, Bin et al. [35] demonstrated that compressive stress–strain diagrams generally
raised as the pore sizes increased. The authors, however, speculated that this was “related to a
change in aspect ratio of the wall thickness against the edge length” [35]. Unlike the aforementioned
research works wherein increasing pore size resulted in either increasing or decreasing the compression

101
Metals 2017, 7, 202

strength, in the study of Xu et al. on biomedical porous NiTi alloys, [36] the strength values varied
with pore sizes in a “S type” fashion, i.e., as the pore size increased once, the value of the strength
dropped initially, and after the second and third increase of the pore size, the strength increased and
decreased, respectively. They attributed this behavior of porous NiTi alloys to the pore size as well
as the number of pores both of which represent the same effect on the mechanical properties of the
porous alloys, i.e., the increase of the pore size and the number of pores would result in a decrease
in the value of mechanical properties including the Rockwell hardness, compression strength and
elastic modulus. The similar pattern in variation of the mechanical properties with increasing pore
size was observed in this study (V type). According to the information provided by the manufacturer,
the nominal pore density of iron foams with nominal cell size of 450, 580, 800 μm is 100–110, 90–100,
and 60–70 ppi (the number of pores per linear one inch). Therefore, it is expected that the number of
pores in IF80 specimens to be significantly lower than that of IF45 and IF58 specimens around 38%
and 32%, respectively. On the other hand, increasing the average pore size from IF58 to IF80 is around
4% for both pore size ranges of 100 < pore size < 200 and pore size > 200 (using the data provided
in Table S1). Therefore, it is speculated that the decrease of the number of pores is more influential
than the increase of the pore sizes, helping to raise the compression strength from IF58 to IF80 sample.
The variation of energy absorbability per volume up to densification strain (W) had a direct relationship
with compression strengths. This is expected given the behavior of cellular materials in plateau regime,
i.e., an increase in the compression strength is associated with the rise of plateau region which results
in increasing the area underneath the stress–strain curve. A linear relationship between the energy
absorption capacity and plastic collapse stress (compressive strength) of open-cell 6101 aluminum alloy
foams under compression was also observed in the work of Krishna et al. [37]. The same V-type pattern
was also observed for variation of the offset yield strength and densification strain values. However,
for the case of elastic modulus values under dry compressions, although the V-type variation was still
present, an important difference was noticed: IF80 specimens showed the highest values of E unlike
the other four properties (σy , σc , εD and W) for which IF45 specimens showed the maximum values.
This can be due to the high influence of the end and branch strut thicknesses which are significantly
higher for IF80 specimens than those of the other two groups (Table S1). The effect of cell size and
relative density on the elastic behavior of an open-cell polymer foam was studied by Maheo et al. [38].
Their experimental results showed that the increase of cell size and relative density led to the increase
of the elastic modulus of the foams.
Given that multiple factors potentially control the mechanical behavior of the iron foams,
predicting the mechanical behavior of the foams under compression is a complex task. Therefore,
a simple tabular method which takes into account multiple parameters, i.e., cell size, pore size, number
of pores and strut thicknesses is proposed to explain the V-shape variation of compression strengths of
the foams studied in this work. Nonetheless, it should be noted that some of the assumptions may
be simplistic. In addition, the method is only proposed as an initial step to develop a model which
predicts the variation of the mechanical properties of open cell foams (compression strength for the
case of this study) as a function of the geometrical properties. Thus, a more sophisticated model may
be developed in the future works. The assumptions of the tabular method, based on the literature
review, are as follow:

- The compression strength has a direct linear relationship with relative density and strut thickness,
and it has an inverse relationship with the pore sizes and the number of the pores.
- All the considered properties have an equal weight to affect the compression strength of the foams.

In the proposed method, for each geometrical property, e.g., relative density, an effect value
(EV) which represents the effect of the parameter on the compression strength level of a particular
group of open-cell foam is determined. The EV is assigned to each group of iron foams (maximum
EV = 100) depending on the average value of the property in the corresponding sample (IF45, IF58,
IF80). For each property, the maximum EV of 100 is assigned to the group which would obtain the

102
Metals 2017, 7, 202

maximum compression strength, if the corresponding property was the only factor influencing the
mechanical strength. For example, for the case of relative density, IF45 sample receive EV of 100,
because it has the maximum relative density among all the samples. Therefore, IF45 specimens would
have the maximum strength among other cell-size group, if the only determining factor was relative
density. Alternatively, for the case of the pore-number factor, IF80 sample is given EV of 100, for IF80
specimens have the lowest number of pores (according to the ppi data). Therefore, if the number of
pore was the only parameter to determine the compression strength, IF80 specimens would have the
highest strength among all the sample groups of iron foams. Then, after assigning the maximum point
for a sample, the EV of the other groups were determined based on their value of the corresponding
property relative to that of the sample with maximum EV. For example, for the case of relative density,
the EV of the IF58 and IF80 samples was 71.05 and 65.79, respectively, which are resulted from the
following equations:
(ρ ∗ /ρS ) IF58 0.027
100 × = 100 × = 71 (2)
(ρ ∗ /ρS ) IF45 0.038
(ρ ∗ /ρS ) IF80 0.025
100 × = 100 × = 66 (3)
(ρ ∗ /ρS ) IF45 0.038
More details on the calculations of the points are available in Section C of the Supplementary
Materials. All the assigned points are shown in Table 4. Considering the described EV assignment
process, a sample with higher total average EV is expected to have a higher compression strength and
therefore energy absorbability. Please note that the final assigned EV associated with pore size are the
result of taking the average of the EVs for pore sizes larger than 200 and those of between 100 and
200 μm using the data of Table S1. Similarly, the assigned EVs for each group with respect to the strut
thickness were the average of the EVs calculated for end-strut and branch-strut thicknesses. That is
why a maximum EV of 100 is not shown in those columns. As expected, IF45 and IF58 groups have
the highest and the lowest average total EV, respectively (V-type variation).

Table 4. Assigned EVs of each sample group for the corresponding geometrical properties.

Specimen ρ ∗ /ρS Cell Size Pore Size Number of Pores Strut Thickness Average Total Point
IF45 100 100 98.46 61.9 72.52 86.43
IF58 71.05 74.75 87.27 68.42 69.54 74.21
IF80 65.79 55.76 83.66 100 100 81.04

4.2. Effect of Environmental Conditions on Compression Behavior of Iron Foams


The statistical analysis of the mechanical responses under wet and dry compression tests suggests
that the wet environment did not significantly influence the mechanical behavior of the iron foams.
This could be due to the presence of open cells which let the fluid, which had a low viscosity, to escape
as the compression applied. Moreover, because the compression was applied shortly after immersion
of the specimens in Hanks’ solution, no degradation effect is expected. Under the condition of higher
strain rate and presence of a more viscous fluid inside the cells, an increase in the strength of the foams
would have been expected: As an open cell containing a fluid is compressed, more work is needed to
act against the viscosity, and a faster deformation of the foam requires more work [28]. However, for
the case of this study, the strain rate and the viscosity were so low that their effects on the mechanical
properties were negligible.
Comparing the compression behavior of IF80 non-immersed specimens with those of the
specimens which were immersed for three- and seven-days showed that the elastic modulus, yield and
compression strengths and energy absorption significantly decreased after immersion. This may be
explained by degradation of iron due to the corrosion, resulting in lowering the level of the mechanical
properties. However, the differences between the three-day-immersed and seven-day-immersed
samples were not statistically significant which can be justified by decreasing the degradation rate

103
Metals 2017, 7, 202

after a few days. Such degradation behavior was observed in some of the pure iron samples after
10- or 15-day dynamic immersion test in Hanks’ solution in the work of Mariot et al. [39]. This was
potentially attributed to the formation of corrosion products on the surface such as iron phosphate
which impeded additional degradation “by hindering oxygen diffusion” [39]. In addition, in [40],
presence of hydroxide layer on the surface of electroformed iron was pointed as a cause of slowing
down the degradation process in the static immersion test. As it appears in Figure 9, layers of corrosion
products formed on the struts after seven-days are thicker than those formed after three-day immersion.
Thus, this may have contributed to the reduction of corrosion rate.

4.3. Deformation and Failure Mechanism


As pointed out previously, three regimes exist during compression of foam: linear elastic, plateau
and densification. In open cell foams with low relative densities (ρ*/ρs ≤ 0.1), cell wall bending mainly
controls the linear elastic regime. Plastic collapse of the cells during compression gives rise to the
plateau of the metal foams which results in formation of plastic hinges in the cell structure. It should
be noted that formation of plastic hinges during plateau occurs in the foams made by materials that
experience plastic yielding such as metals or rigid polymers [28]. Plastic hinges were formed in iron
foam specimens as shown in Figure 8a. Finally, densification takes place “when the cells have almost
completely collapsed” and further strains require much higher level of stress when cell walls being
adjacent to each other [28]. Figure 10 represents a simple model of an open cell foam under linear
elastic deformation and the formation of plastic hinges during plastic collapse.

Figure 10. A simple model of an open cell foam experiencing: (a) linear elastic deformation; and (b)
formation of plastic hinges during plastic collapse. Adapted from [28].

Different modes of plastic deformation of struts, including S-shape and C-shape deformations nd
deformation bands in IF45 specimens are shown in Figure 8. Although Gibson’s report [29] noted that
the buckling of cells takes place during softening of aluminum foams, in this study, the presence or
absence of buckling cannot be verified with a high level of certainty as the SEM images show only
the surfaces on which compression applied. In other words, the struts whose directions are along the
loading axis are not clearly visible. However, given that many of the struts are inclined to the loading
axis, it is expected that buckling partially contributes to many of the strut deformatins. In Kaya and
Fleck’s work [27], both bending and buckling were addressed as two different deformation mechanisms
appeared in different struts. In Daxner’s work [41], both buckling and bending are pointed out as
“dominating deformation mechanism” in the struts of “open-cell metallic foams”. In the work of
Schuler et al. [26], it is stated that C-shape and S-shape deformations are the product of bending and
torsion. Therefore, given that the strut orientation relative to the loading direction is an influential
factor on deformation behavior of the foams [37], it is fair to state that the deformation of an open
cell iron foam under compression is a complex mechanism which could be the product of different
mechanisms in conjunction with each other such as bending, buckling and torsion. However, some of

104
Metals 2017, 7, 202

the struts may experience only one form of deformation. The complexity of the strut deformations has
been acknowledged in [26] as well.

5. Conclusions
To investigate the influence of geometrical properties and environmental condition on mechanical
behavior of porous iron, iron foams of three different cell sizes underwent dry and wet compression
tests. In addition, mechanical properties of iron foam specimens after static immersion were assessed
and compared with those of non-immersed specimens. Using the stress–strain responses obtained
from compression tests, elastic and plastic mechanical properties were analyzed: elastic modulus,
yield strength, compression strength, densification strain and energy absorbability per volume up
to the point of densification. Two major groups of statistical tests were carried out to analyze the
significance of difference between the values of properties: hypothesis tests including ANOVA in
conjunction with Tukey post hoc test and t-test, and non-parametric tests including Kruskal–Wallis
ANOVA and two-sample K–S tests. In most, though not all, cases, hypothesis and non-parametric tests
led to similar conclusions.
It was observed that the mechanical properties of the foams under dry compression tests had
a “V-type” variation. Comparing the values of elastic modulus revealed that IF80 specimens had
a the highest level of stiffness while for other properties, i.e., yield strength, compression strength,
densification strain and energy absorbability, IF45 specimens possessed the highest level under dry
compression. A simple tabular method was proposed to explain the variation in compression strength
of the iron foams of different geometrical properties with respect to each other. Wet environment
generally did not alter the mechanical behavior of the iron foams significantly while degradation
decreased the elastic modulus, yield and compression strength and the energy absorbability of the
samples. The deformation behavior of the foams under compression was also evaluated via SEM
images and different deformation modes were identified. It was speculated that the deformation
of open cell iron foams under compression is a complex phenomenon that could be the product of
multiple mechanism such as bending, buckling and torsion. However, further studies are needed to
understand the failure mechanism of iron foams.

Supplementary Materials: The following are available online at http://www.mdpi.com/2075-4701/7/6/202/s1.


Figure S1: Measurement of cell size, pore size and different strut thicknesses on an IF58 specimen, Figure S2:
Experimental design and parameters with the number of specimens used in each group, Figure S3: Stress-strain
curves of IF45 samples under compression: (a) compression area A1 and cross-head speed S1 , (b) compression
area A2 and cross-head speed S1 , (c) compression area A1 and cross-head speed S2 , (d) compression area A2
and cross-head speed S2 , Figure S4: Stress-strain curves of IF58 samples under compression: (a) compression
area A1 and cross-head speed S1 , (b) compression area A2 and cross-head speed S1 , (c) compression area A1 and
cross-head speed S2 , (d) compression area A2 and cross-head speed S2 , Figure S5: Stress-strain curves of IF80
samples under compression: (a) compression area A1 and cross-head speed S1 , (b) compression area A1 and
cross-head speed S2 , (c) compression area A2 and cross-head speed S1 , (d) compression area A2 and cross-head
speed S2 , Table S1: Results of the measurement.
Acknowledgments: This work is supported by the Natural Sciences and Engineering Research Council of Canada
(NSERC) via the Discovery Grant.
Author Contributions: Hendra Hermawan and Adhitya Trenggono conceived and designed the experiments;
Reza Alavi, Adhitya Trenggono and Sébastien Champagne performed the experiments; Reza Alavi and Hendra
Hermawan analyzed the data; Reza Alavi wrote the paper; Hendra Hermawan revised and approved the paper.
Conflicts of Interest: The authors declare no conflict of interest.

References and Note


1. Pippan, R.; Motz, C.; Kriszt, B.; Zettl, B.; Mayer, H.; Stanzl-Tschegg, S.; Simancik, F.; Kovacik, J. Material
Properties. In Handbook of Cellular Metals: Production, Processing, Applications; Degischer, H.P., Kriszt, B., Eds.;
Wiley-VCH: New York, NY, USA, 2002.
2. Ashby, M.F.; Evans, A.G.; Fleck, N.A.; Gibson, L.J.; Hutchinson, J.W.; Wadley, H.N.G. Properties of Metal
Foams in Metal Foams: A Design Guide; Butterworth-Heinemann: Oxford, UK, 2000; pp. 40–54.

105
Metals 2017, 7, 202

3. Wu, S.; Liu, X.; Yeung, K.W.K.; Liu, C.; Yang, X. Biomimetic porous scaffolds for bone tissue engineering.
Mater. Sci. Eng. R Rep. 2014, 80, 1–36. [CrossRef]
4. He, J.; He, F.L.; Li, D.W.; Liu, Y.L.; Yin, D.C. A novel porous Fe/Fe–W alloy scaffold with a double-layer
structured skeleton: Preparation, in vitro degradability and biocompatibility. Colloids Surf. B Biointerfaces
2016, 142, 325–333. [CrossRef] [PubMed]
5. Alvarez, K.; Nakajima, H. Metallic scaffolds for bone regeneration. Materials 2009, 2, 790–832. [CrossRef]
6. Wen, Z.; Zhang, L.; Chen, C.; Liu, Y.; Wu, C.; Dai, C. A construction of novel iron-foam-based calcium
phosphate/chitosan coating biodegradable scaffold material. Mater. Sci. Eng. C 2013, 33, 1022–1031.
[CrossRef] [PubMed]
7. Čapek, J.; Vojtěch, D.; Oborná, A. Microstructural and mechanical properties of biodegradable iron foam
prepared by powder metallurgy. Mater. Des. 2015, 83, 468–482. [CrossRef]
8. Eliaz, N. Biodegradable Metals. In Degradation of Implant Materials; Springer Science and Business Media:
New York, NY, USA, 2012; p. 94.
9. Peuster, M.; Wohlsein, P.; Brügmann, M.; Ehlerding, M.; Seidler, K.; Fink, C.; Brauer, H.; Fischer, A.;
Hausdorf, G. A novel approach to temporary stenting: Degradable cardiovascular stents produced from
corrodible metal-results 6–18 months after implantation into New Zealand white rabbits. Heart 2001, 86,
563–569. [CrossRef] [PubMed]
10. Oriňáková, R.; Oriňák, A.; Bučková, L.M.; Giretová, M.; Medvecký, L.; Labbanczová, E.; Kupková, M.;
Hrubovčáková, M.; Kova, M. Iron based degradable foam structures for potential orthopedic applications.
Int. J. Electrochem. Sci. 2013, 8, 12451–12465.
11. Zberg, B.; Uggowitzer, P.J.; Löffler, J.F. MgZnCa glasses without clinically observable hydrogen evolution for
biodegradable implants. Nat. Mater. 2009, 8, 887–891. [CrossRef] [PubMed]
12. Vojtěch, D.; Kubásek, J.; Šerák, J.; Novák, P. Mechanical and corrosion properties of newly developed
biodegradable Zn–based alloys for bone fixation. Acta Biomater. 2011, 7, 3515–3522. [CrossRef] [PubMed]
13. Vojtěch, D.; Kubásek, J.; Čapek, J.; Pospíšilová, I. Comparative mechanical and corrosion studies on
magnesium, zinc and iron alloys as biodegradable metals. Mater. Technol. 2015, 49, 877–882. [CrossRef]
14. Kraus, T.; Moszner, F.; Fischerauer, S.; Fiedler, M.; Martinelli, E.; Eichler, J.; Witte, F.; Willbold, E.;
Schinhammer, M.; Meischel, M. Biodegradable Fe–based alloys for use in osteosynthesis: Outcome of
an in vivo study after 52 weeks. Acta Biomater. 2014, 10, 3346–3353. [CrossRef] [PubMed]
15. Hermawan, H.; Alamdari, H.; Mantovani, D.; Dubé, D. Iron–manganese: New class of metallic degradable
biomaterials prepared by powder metallurgy. Powder Metall. 2008, 51, 38–45. [CrossRef]
16. Yusop, A.H.M.; Daud, N.M.; Nur, H.; Kadir, M.R.A.; Hermawan, H. Controlling the degradation kinetics of
porous iron by poly(lactic-co-glycolic acid) infiltration for use as temporary medical implants. Sci. Rep. 2015,
5, 11194. [CrossRef] [PubMed]
17. Hermawan, H.; Dubé, D.; Mantovani, D. Degradable metallic biomaterials: Design and development of
Fe–Mn alloys for stents. J. Biomed. Mater. Res. Part A 2010, 93, 1–11. [CrossRef] [PubMed]
18. Oh, K.; Lee, E.; Bae, J.S.; Jang, M.J.; Poss, R.; Kieback, B.; Walther, G.; Kloeden, B. Large scale production and
applications of alloy metal foam. Metfoam 2011 Proc. 2011, S, 601–606.
19. Gibson, L.J.; Ashby, M.F. The structure of cellular solids. In Cellular Solids: Structure and Properties, 2nd ed.;
Cambridge University Press: Cambridge, UK, 1997.
20. International Organization for Standardization. ISO13314: 2011(E), Mechanical testing of metals_Ductility
testing_Compression test for porous and cellular metals, Switzerland, 2011-12-15.
21. Banhart, J.; Baumeister, J. Deformation characteristics of metal foams. J. Mater. Sci. 1998, 33, 1431–1440.
[CrossRef]
22. Paul, A.; Ramamurty, U. Strain rate sensitivity of a closed-cell aluminum foam. Mater. Sci. Eng. A 2000, 281,
1–7. [CrossRef]
23. Beer, F.P.; Johnston, E.R.; Dewolf, J.T.; Mazurek, D.F. Mechanics of Materials, 6th ed.; McGraw-Hill: New York,
NY, USA, 2012.
24. Montanini, R. Measurement of strain rate sensitivity of aluminium foams for energy dissipation. Int. J.
Mech. Sci. 2005, 47, 26–42. [CrossRef]
25. OriginLab. Nonparametric Tests. Available online: http://www.originlab.com/index.aspx?go=Products/
Origin/Statistics#Nonparametric_Tests_PRO (accessed on 31 March 2017).

106
Metals 2017, 7, 202

26. Schüler, P.; Fischer, S.F.; Bührig-Polaczek, A.; Fleck, C. Deformation and failure behaviour of open cell Al
foams under quasistatic and impact loading. Mater. Sci. Eng. A 2013, 587, 250–261.
27. Kaya, A.C.; Fleck, C. Deformation behavior of open-cell stainless steel foams. Mater. Sci. Eng. A 2014, 615,
447–456. [CrossRef]
28. Gibson, L.J.; Ashby, M.F. The mechanics of foams: Basic results. In Cellular Solids: Structure and Properties,
2nd ed.; Cambridge University Press: Cambridge, UK, 1997.
29. Gibson, L.J. Mechanical behavior of metallic foams. Annu. Rev. Mater. Sci. 2000, 30, 191–227. [CrossRef]
30. Amsterdam, E.; Vries, J.H.B.; Hosson, J.T.M.; Onck, P.R. The influence of strain-induced damage on the
mechanical response of open-cell aluminum foam. Acta Mater. 2008, 56, 609–618. [CrossRef]
31. Nieh, T.G.; Higashi, K.; Wadsworth, J. Effect of cell morphology on the compressive properties of open-cell
aluminum foams. Mater. Sci. Eng. A 2000, 283, 105–110. [CrossRef]
32. Stephani, G.; Andersen, O.; Göhler, H.; Kostmann, G.; Kümmel, K.; Quadbeck, P.; Reinfried, M.; Studnitzky, T.;
Waag, U. Iron Based Cellular Structures—Status and Prospects. Adv. Eng. Mater. 2006, 8, 847–852. [CrossRef]
33. Xiao, L.; Song, W.; Wang, C.; Liu, H.; Tang, H.; Wang, J. Mechanical behavior of open-cell rhombic
dodecahedron Ti–6Al–4V lattice structure. Mater. Sci. Eng. A 2015, 640, 375–384. [CrossRef]
34. Jian, Y.T.; Yang, Y.; Tian, T.; Stanford, C.; Zhang, X.P.; Zhao, K. Effect of Pore Size and Porosity on the
Biomechanical Properties and Cytocompatibility of Porous NiTi Alloys. PLoS ONE 2015, 10, e0128138.
[CrossRef] [PubMed]
35. Jiang, B.; Wang, Z.; Zhao, N. Effect of pore size and relative density on the mechanical properties of open cell
aluminum foams. Scr. Mater. 2007, 56, 169–172. [CrossRef]
36. Xu, J.L.; Bao, L.Z.; Liu, A.H.; Jin, X.F.; Luo, J.M.; Zhong, Z.C.; Zheng, Y.F. Effect of pore sizes on the
microstructure and properties of the biomedical porous NiTi alloys prepared by microwave sintering.
J. Alloys Compd. 2015, 645, 137–142. [CrossRef]
37. Krishna, B.V.; Bose, S.; Bandyopadhyay, A. Strength of open-cell 6101 aluminum foams under free and
constrained compression. Mater. Sci. Eng. A 2007, 452–453, 178–188. [CrossRef]
38. Maheo, L.; Viot, P.; Bernard, D.; Chirazi, A.; Ceglia, G.; Schmitt, V.; Mondain-Monval, O. Elastic behavior of
multi-scale, open-cell foams. Compos. Part B Eng. 2013, 44, 172–183. [CrossRef]
39. Mariot, P.; Leeflang, M.A.; Schaeffer, L.; Zhou, J. An investigation on the properties of injection-molded pure
iron potentially for biodegradable stent application. Powder Technol. 2016, 294, 226–235. [CrossRef]
40. Moravej, M.; Purnama, A.; Fiset, M.; Couet, J.; Mantovani, D. Electroformed pure iron as a new biomaterial
for degradable stents: In vitro degradation and preliminary cell viability studies. Acta Biomater. 2010, 6,
1843–1851. [CrossRef] [PubMed]
41. Daxner, T. Deformation Mechanisms and Yielding in Cellular Metals. In Plasticity of Pressure-Sensitive
Materials; Altenbach, H., Öchsner, A., Eds.; Springer: Berlin/Heidelberg, Germany, 2014; pp. 165–166.

© 2017 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

107
metals
Review
Biodegradable Metallic Wires in Dental and
Orthopedic Applications: A Review
Mohammad Asgari 1,2 , Ruiqiang Hang 1,3 , Chang Wang 4,5 , Zhentao Yu 4,5 , Zhiyong Li 1,2, * and
Yin Xiao 1,2,6, *
1 The Institute of Health and Biomedical Innovation, Queensland University of Technology,
Brisbane, QLD 4059, Australia; m.asgari@hdr.qut.edu.au
2 School of Chemistry, Physics & Mechanical Engineering, Science & Engineering Faculty,
Queensland University of Technology, Brisbane, QLD 4000, Australia
3 Research Institute of Surface Engineering, Taiyuan University of Technology, Taiyuan 030024, China;
hangruiqiang@tyut.edu.cn
4 Northwest Institute for Nonferrous Metal Research, Shaanxi Key Laboratory of Biomedical Metal Materials,
Xi’an 710016, China; cch_wang@163.com (C.W.); yzt@c-nin.com (Z.Y.)
5 China-Australia Joint Research Centre of Biomedical Metallic Materials, Shaanxi Key Laboratory of
Biomedical Metal Materials, Xi’an 710016, China
6 The Australia-China Centre for Tissue Engineering and Regenerative Medicine (ACCTERM),
Queensland University of Technology, Brisbane, QLD 4059, Australia
* Correspondence: zhiyong.li@qut.edu.au (Z.L.); yin.xiao@qut.edu.au (Y.X.); Tel.: +61-7-3138-5112 (Z.L.);
+61-7-3138-6240 (Y.X.); Fax: +61-7-3138-8381 (Z.L.); +61-7-3138-6030 (Y.X.)

Received: 30 January 2018; Accepted: 22 March 2018; Published: 26 March 2018

Abstract: Owing to significant advantages of bioactivity and biodegradability, biodegradable metallic


materials such as magnesium, iron, and zinc and their alloys have been widely studied over recent
years. Metallic wires with superior tensile strength and proper ductility can be fabricated by a
traditional metalworking process (drawing). Drawn biodegradable metallic wires are popular
biodegradable materials, which are promising in different clinical applications such as orthopedic
fixation, surgical staples, cardiovascular stents, and aneurysm occlusion. This paper presents
recent advances associated with the application of biodegradable metallic wires used in dental
and orthopedic fields. Furthermore, the effects of some parameters such as the surface modification,
alloying elements, and fabrication process affecting the degradation rate as well as biocompatibility,
bioactivity, and mechanical stability are reviewed in the most recent works pertaining to these
materials. Finally, possible pathways for future studies regarding the production of more efficient
biodegradable metallic wires in the regeneration of bone defects are also proposed.

Keywords: magnesium; zinc; iron; biodegradable materials; wire drawing; bone regeneration;
bone tissue engineering

1. Introduction
Metallic wires have been widely used in a wide range of industries in different applications.
Drawing is a conventional metal forming process which has been applied to fabricate metallic wire.
Briefly, this process continuously decreases the cross-sectional area of a specimen by pulling it through
a single (in single-pass drawing) or series (in the multi-pass drawing) of the conical die(s) [1]. Figure 1
shows the schematic of die geometry in single-pass wire drawing process. di , do and α are the input
diameter, the output diameter and die angle respectively. Due to imparting of a huge cold working
happened during the drawing process, production of thin metallic wire with considerable strength is

Metals 2018, 8, 212; doi:10.3390/met8040212 108 www.mdpi.com/journal/metals


Metals 2018, 8, 212

potentially achievable [2]. Equation (1) represents the percentage of cold work that could be achieved
by single-pass wire drawing process [3]:
 2
do
Cold work (%) = 1 − (1)
di

Figure 1. Schematic representation of a typical pass in the wire drawing process.

Moreover, different intermediate and/or post heat treatments coinciding with precise die(s)
design and manufacturing could lead to the fabrication of the fine wire with acceptable ductility for
different applications [3].
Over recent decades, the application of the wire drawing process in the production of biomedical
metallic wires has been investigated. Drawn wires from titanium (Ti) and its alloys, stainless steels (SSs),
and nitinol have been used in different medical devices such as cardiovascular stents [4], coronary guide
catheter [5], coil occlusion of the aneurysm [6], Kirschner wires (K-wire) [7], orthodontic archwires [8],
ligature wire [9], surgical sutures and versus staples [10]. Based on the application, such devices might
be implanted either for a short or long time in the human body and regardless of the duration time,
after finishing their mission, they should be removed from the body by a secondary operation. Because
they cannot be degraded by human body fluids and, in some cases, if they are kept for more time, they
will lead to complications such as harsh inflammations [11].
To decrease the clinical expenses and increase the quality of life of patients by reducing any
secondary surgeries for removal of metal implants as well as minimizing any negative consequences,
a new generation of metallic biomaterials has been introduced as a temporary support. These kinds
of materials, called biodegradable metals (BMs), are strong enough during the target tissue healing
process and can be gradually resorbed in the human body [12,13]. The degradation of BMs mainly
ascribes to electrochemical corrosion, during which process the metal atoms lose their free electrons,
become cations and enter body fluids. The content and the releasing allowance of those ions should
not be toxic to the human body, especially for the surrounding host tissues [14]. To that aim, up to now
different biodegradable alloys have been introduced. Among them, magnesium (Mg), iron (Fe), zinc
(Zn), and their alloys as biodegradable implant metals are the three main categories that are considered
more than other beneficial elements such as calcium (Ca) and strontium (Sr) [13].
In addition to the structural role of Mg, Fe, Zn, and their alloys as an implant device, they play
important roles in many metabolic reactions and biological mechanisms of the body and consequently
reduce the healing duration [15]. That is the reason they are named nutrient metals (NMs) [14,16]. Up to
now, many studies have been conducted to increase the performance of these NMs by enhancing their
initial mechanical properties and mitigating their potential drawbacks such as imperfect degradation
rate to convert them from mechanical replacement devices to nutrient implants.

109
Metals 2018, 8, 212

As an implant device, BMs in the shape of wire were firstly applied in 1878 by Edward C. Huse.
He used pure Mg wire as a ligature to stop bleeding vessels. Although Huse reported the first clinical
application of Mg wire, the Austrian surgeon Prof. Dr. Erwin Payr is known as the pioneer in the field
of biodegradable Mg implants. His valuable results and consequently his inspiring articles published
from 1892 to 1905 led to the utilization of biodegradable metallic wires in both human and animal
models by other researchers [17]. Andrews in 1917 [18], and Seelig in 1924 [19] were early followers
who investigated the animal in vivo experiments of pure Mg wire. From that time, a considerable
number of feasibility studies have been done to propose Mg, Fe and more recently Zn and their alloys
as ideal candidates in interventional clinical devices such as cardiovascular stents [20], and orthopedic
implants [1,21]. However, there are still some challenges including improvement of both the surface
and bulk properties such as corrosion resistance, bioactivity, strength, ductility, and toxicity. These
properties need to be modified by more investigations to approve the application of BMs and their
alloys in different medical operations without any negative consequences about the health of patients.
The present review mainly focuses on the recent progress related to the usage of biodegradable
metallic wires in the orthopedic and dental studies. To that aim, the positive and any possible negative
effects of Mg, Fe, and Zn on the bone biological mechanisms and bone healing process are generally
presented first. Then, the history of clinical application development of each of these metals and their
alloys in the shape of wires is reviewed. Furthermore, any other related scientific results to in vitro
and in vivo studies of the biodegradable metallic wires are briefly summarized. Finally, the remaining
problems in this research field that could be addressed in the future work are discussed and some
possible options that are beneficial to the improvement of the clinical performance of these wires in the
bone applications are also proposed.

2. Biological and Mechanical Aspects of Biodegradable Metallic Materials for Bone


Regeneration Applications

2.1. Mg and Its Alloys

2.1.1. Advantages of Mg and Its Alloys


Mg as a mineral is the fourth most abundant cation element in the body [22]. It has a pivotal
role in many body metabolisms including enzymatic reaction, the formation of apatite, and bone cells
adsorption [23]. About half of the total amount of Mg in our body is stored in bone tissue [24]. Wu et al.
reported that Mg ions (Mg2+ ) govern a stimulating effect on the new bone formation by increasing the
proliferation and differentiation of osteogenic (stem) cells via osteogenesis-related signaling pathways
in vitro [25]. The release of Mg2+ is related to the chemical reaction of Mg in the human body fluids
based on the following reaction which produces Mg hydroxide (Mg(OH)2 ) and hydrogen gas (H2 )
evolution [26]:
Mg(s) + 2H2 O(aq) → Mg(OH)2 (s) + H2 ↑ (g) (2)

Zheng et al. [27] presented a degradation mechanism for Mg in the physiological environment by
the following anodic dissolution and cathodic reduction reactions:

Mg(s) → Mg2+ (aq) + 2e− (anodic reaction) (3)

2H2 O(aq) + 2e− → H2 (g) ↑ +2OH− (aq) (cathodic reaction) (4)



Mg 2+
(aq) + 2OH → Mg(OH)2 (s) (cathodic reaction) (5)

It is clear that Mg2+ originates from the anodic reaction. Then, some of them take part in a reaction
with hydroxide ions (OH− ) to produce Mg(OH)2 , while others enter the solution.

110
Metals 2018, 8, 212

Mg(OH)2 , which is called milk of magnesia, includes both water and MgO solid particles. Hence,
some part of it can convert to MgO and vice versa (see Equation (6)):

Mg(OH)2 (aq) ↔ MgO(s) + H2 O(aq) (6)

Both of Mg(OH)2 and MgO are beneficial inorganic materials to the body [28]. In addition
to the conventional medical application of Mg(OH)2 which has been widely used as an antacid to
neutralize stomach acid and a laxative, it also possesses numerous valuable properties including
excellent biocompatibility, nontoxic antibacterial activity, and high drug loading ability [29]. In bone
regeneration application, as a major degradation product from any magnesium alloys, Mg(OH)2 can
enhance the bone formation and temporarily decrease the bone resorption resulting in a higher bone
mass [30,31]. Therefore, it could be a potential protective coating material for Mg-based implants to
reduce the corrosion rate and improve bone ingrowth simultaneously [31].
Regardless of the wide range of industrial applications, MgO similar to Mg(OH)2 is applied
as both an antacid and a laxative. It is also used as a drying agent because it rapidly reacts
with water and creates Mg(OH)2 [28]. MgO in shape of nanoparticles has been introduced as a
novel antibacterial material to enhance tissue regeneration process and reduce bacterial infection
during orthopedic implantations [32]. Huang et al. [33] claimed that the toxicity of MgO against
bacteria increased by reducing its particle size. Hickey et al. [32] fabricated nanocomposites
using MgO nanoparticles to mineralize poly(L-lactic acid) (PLLA) and applied in traumatic brain
injury(TBI)-related orthopedic tissue regeneration. They reported that all the seeded bacteria were
killed by PLLA/MgO nanocomposite. Therefore, MgO, as one of the corrosion products of Mg and its
alloys play an antibacterial role for orthopedic tissue engineering applications.
In terms of mechanical aspects, Mg and its biocompatible alloys are more advantageous in bone
applications in comparison with other metallic materials, ceramics, and biodegradable polymers
(BPs). Because the mechanical properties of Mg are more similar to those of natural cortical bone
compared to other materials which can decrease the stress shielding effect and prevent consequent bone
resorption in the implanted sites [34,35], that makes Mg the most attractive biodegradable material in
the orthopedic devices and implants as a suitable substitution for the non-bioresorbable one. Table 1
summarizes the information about the typical mechanical properties of cortical bone and commonly
used metallic biomaterials.

Table 1. Comparison of typical mechanical properties of cortical bone and commonly used metallic
biomaterials, data from [35–37].

Young’s Modulus, Yield Tensile Ultimate Tensile Density


Tissue/Material
E (GPa) Strength (MPa) Strength (MPa) (g/cm3 )
Cortical bone 5–23 104.9–114.3 35–283 1.8–2.0
Pure Mg, wrought 41–45 100 180 1.74
Pure Zn, as cast 90 10 20 7.13
Pure Fe 211.4 50 540 7.87
Ti-6Al-4V, as cast 114 760–880 830–1025 4.43
316L stainless steel 193 200–300 450–650 8
Co-Cr alloy 240 500–1500 900–1540 8.3

2.1.2. Disadvantages of Mg and Its Alloys


The standard electrode potential for Mg(s) ↔ Mg2+ (aq) + 2e− reaction is very low (−2.37 V) [38],
making Mg very reactive in aqueous solutions such as body fluids. Although there is a gap between
in vitro and in vivo results regarding the corrosion mechanism of Mg [39], commonly for both in vitro
and in vivo the oxide film formed on the surface of Mg is porous rather than dense, thus cannot
effectively prevent direct contact between metallic Mg and solution, leading to the rapid degradation
of Mg and its alloys. The reduction in the mechanical stability due to the stress corrosion cracking is the

111
Metals 2018, 8, 212

first negative consequence related to such high corrosion rate of Mg and its alloys, risking the adequate
load shielding over the tissue regeneration duration, especially in load-bearing applications [13,40].
Therefore, it is vitally important that there should be a balance between the degradation time and
the healing period of damaged tissue to prevent implantation mechanical failure. It is the point that
should be considered in the geometry design, material selection, and fabrication process of all BMs
and their alloys generally, and more critical for Mg and its alloys.
Hydrogen gas (H2 ) evolution is another challenging issue in the application of Mg and its alloys
in the body. Based on Equation (2), 1 g of pure Mg could produce 1 liter (L) of H2 gas [41]. Rely on
previous in vivo experiments, some of this gas is soluble in the blood. The solubility of H2 depends
on the amount of the blood flow in the implantation site. The remained H2 bubbles are accumulated
in the local host tissue and create gas cavity endangering the cell adhesion to implants [42]. In some
applications such as cardiovascular stents, due to the convective transport phenomena, the evolution
of H2 is not as concerning as orthopedic applications; because bone inherently is a poorly vascularized
tissue leading to the formation of H2 cavities potentially [26]. Some animal studies have shown
that it might be dangerous [43]. In 2016, Noviana et al. [41] investigated the effect of H2 evolution
from porous pure-magnesium implants on the mortality of adult rats. They reported that the cavity
formation was not tolerable for all the rats and the survival rate at Day 18 post-implantation reached
zero percent (see Figure 2). Hence, such amount of post-implantation mortality is a warning sign for
the risk of excessive H2 accumulation in the large implants such as fixation plates. Nonetheless, due to
the very light mass and negligible surface area of fine magnesium wire having sub-millimeter diameter
there should not be that much of concern regarding the risk of H2 gas evolution. However, it is still an
estimation that needs to be investigated precisely by more in vitro and in vivo studies.

Figure 2. (a) Porous pure-magnesium implantation; (b) survival rate of the group of rats with
magnesium implants compared with control group; (c) gas cavity formation; and (d) subcutaneous
opening of the implantation area at Day 7, reproduced with permission from [41], Elsevier, 2016.

2.2. Fe and Its Alloys

2.2.1. Advantages of Fe and Its Alloys


Iron (Fe) as the fourth most abundant element on earth is significantly necessary for the human
body and takes part in many enzymatic systems. The hemoglobin uses the largest part of total Fe

112
Metals 2018, 8, 212

(about 60–70%). Whereas, only about 10% of it is stored in bone marrow [44]. Fe has an essential
role in the function of red blood cells and oxygen transport throughout the body. Fe deficiency may
decrease the number of healthy oxygen-carrying red blood cells, thus has a negative effect on physical
activity and increases susceptibility to infections [45]. Fe has a crucial role in bone metabolism by
participating in two main biological processes including collagen synthesis and vitamin D activation
and deactivation [46]. The clinical observations have shown that both Fe overload and deficiency (with
or without anemia) could happen for the patients who suffered from osteoporosis [45,46]. The link
between Fe and bone metabolism, especially in the homeostasis of bone ageing to maintain a balance
between bone formation and resorption, is still in a need of further investigation, especially for elderly
people to increase the quality of their life.
Almost all of the previous in vivo studies in animal models have been mainly focused on
the application of Fe and its alloys as suitable degradable implant materials in the fabrication of
cardiovascular stents [47]. While there are few studies in literature related to the bone application of Fe
and its alloys. Kraus et al. [48] followed up the degradation behavior of Fe-based wires implanted into
the femur of 38 rats for 52 weeks. Regardless of some results [48] showing that Fe ions could release
from the implanted wires into the surrounding tissues, there is no clear information about its benefits
on bone regeneration process.

2.2.2. Disadvantages of Fe and Its Alloys


The standard electrode potential for the reaction of Fe(s) ↔ Fe2+ (aq) + 2e− is relatively high
(−0.444 V) [49]. Furthermore, owing to including a high concentration of CO2 gas in blood causing to
create a uniform, compact and dense manganese carbonate crystals the degradation rate of Fe in vivo
is slower than in vitro [50] and in most of the cases is longer than clinical needs [51]. Additionally,
because of the low rate of blood circulation near the implants, utilization of Fe and its alloys for
bone reconstruction purposes would be definitely lower than their corrosion rate. Adding some
alloying elements to pure Fe has been reported to be one of the most effective methods to accelerate
the degradation rate of Fe to an acceptable level for practical clinical usage [12,44]. Among all of the
Fe-based alloys, Fe–Mn alloys have been proved to be the most promising ones in accelerating the
corrosion rate of Fe with similar mechanical properties (i.e., Young’s modulus and ultimate tensile
strength) to 316L SS [51]. Moreover, Mn can play an important role in promoting the growth of new
bone and connective tissues as well as reducing bone loss [52,53] which could increase its bioactivity
in orthopedic and dental applications. However, Fe and its alloys have much higher mechanical
properties than those of the human cortical bone (see Table 1), which keeps the stress shielding effect
as a challenging issue to select them as a suitable candidate in the load-bearing locations, especially as
bulk temporary implants.

2.3. Zn and Its Alloys

2.3.1. Advantages of Zn and Its Alloys


Zinc (Zn) as the second most abundant transition metal element in the human body is found
in all organs, tissues, fluids, and body secretions [54,55]. It is considered to be one of the vital
mineral elements participating in a variety of fundamental biological functions, including signal
transduction, apoptosis regulation as well as nucleic acid metabolism, DNA and RNA polymerase, and
organic ligands interactions [56–58]. Approximately 85% of the Zn content of body exists in bone and
muscle [14] which serves a crucial role in the improvement of cell proliferation and osteo-related gene
expressions of osteoblasts affecting bone in growth [59,60]. Li et al. [56] investigated the biological
effects of the Zn-1X(Mg, Ca and Sr) binary alloys using thin pin shape implants inserted into the
bone tunnel created along the axis of the femoral shaft from the distal femur of mice. They reported
that the Zn-1X binary alloys with nutrient alloying elements improved new bone formation around
the pins effectively. Other in vivo and in vitro studies have shown that Zn affects the osteoclastic

113
Metals 2018, 8, 212

bone resorption metabolism [61,62]. The Zn content in the bone matrix of patients suffering from
skeletal diseases such as osteoporosis (commonly in elderly people) and bone cancer is less than that
of normal people [56,63]. In summary, having a healthy bone with sufficient density is correlated with
the balancing of Zn content in the human body.
It has been well established that Zn ions (Zn2+ ) show antibacterial property [64–66]. Recently,
Zhu et al. [59] introduced plasma immersion ion implantation (PIII) process as a promising method
for Zn-incorporation on Ti alloy, which is beneficial to both osteogenic and antibacterial abilities to
enhance osseointegration while decreases implant-associated infections. Zn2+ are released directly
from the corrosion of Zn and its possible alloys when in contact with the human body fluids through
electrochemical corrosion. Drelich et al. proposed [67] the hypothetical mechanism happening during
the degradation of Zn in the physiological environment (Figure 3). The reaction between some of Zn2+
with hydroxyl ions (OH− ) forms Zn oxide (ZnO). In addition to the wide range of dental and orthopedic
applications of ZnO, due to its desirable antibacterial and osteogenic abilities, the anti-cancer property
of ZnO has been also approved as a beneficial material to cure osteosarcoma [68,69].

Figure 3. Schematic diagram of degradation mechanism of Zn in a physiological environment.

The standard potential for the cathodic reaction of Zn is −0.762 V [49], which is between that
of Mg and Fe. Therefore, the moderate degradation rate of Zn is faster than Fe but slower than Mg,
which could be an ideal option for clinical applications, especially in terms of mechanical stability [55].
More recently, Yang et al. [57] implanted pure Zn stents into the abdominal aorta of rabbits for
12 months. They reported that the stent was capable to serve mechanical stability for the first 6 months
and degraded 41.75 ± 29.72% of the stent volume after 12 months of implantation without any
severe inflammations, platelet aggregation, and thrombosis formation. Furthermore, in the matter of
accumulation of H2 during the degradation, there is less concern in the application of Zn and its alloys
in bone applications than that of Mg and its alloys. Because it seems that the corrosion rate of Zn is low
enough to prepare a time for the dissolving of H2 in the body without the creation of any gas cavities.

2.3.2. Disadvantages of Zn and Its Alloys


Generally, pure Zn is soft, brittle with low strength (Table 1). Therefore, it seems impossible to
apply pure zinc in load-bearing applications. The addition of alloying elements has been recognized
as one of the most effective methods to improve the mechanical properties of pure Zn [56]. Even

114
Metals 2018, 8, 212

adding a little amount of alloying elements can improve its mechanical properties considerably. Briefly,
depending on the composition of the Zn alloys the ultimate tensile strength (UTS) has been achieved
in a wide range of 87 to 399 MPa till now [55,70]. Moreover, other parameters such as using different
metalworking processes such as rolling, extortion, and post-processing heat treatments can definitely
affect the strength, ductility, and degradation behavior of Zn-based alloys. However, Young’s modulus
(E) of Zn is about twice as that of Mg, which indicates that the stress shielding effect is more challenging
in Zn-based implants.
Non-uniform corrosion behavior of Zn compared to uniform corrosion of both Mg and Fe is
another problem. Chen et al. [15] compared the corrosion behavior of Zn, Fe and Mg in a long-term
course in vitro. They observed that by increasing immersion time the corrosion rate of Zn developed
faster than Fe and Mg, which is justified by the occurrence of localized corrosion in the surface of
Zn samples. Similar behavior was reported in vivo [57]. The localized corrosion causes to lose the
strength uniformity of an implant which makes its failure behavior more ambiguous. Drelich et al. [67]
claimed that coating of Zn with a passive layer of ZnO film could be an effective approach to delay the
degradation rate and decrease the corrosion non-uniformity. However, Törne et al. [71] claimed that
the corrosion behavior of Zn is affected by the types of examined solution. They proposed that whole
blood or real plasma are suitable solutions for short-term in vitro studies which can resemble in vivo
situation, while Ringer’s solution is more suitable for long-term degradation studies. Collectively,
more in vivo and in vitro investigations should be conducted for the better understanding of the
corrosion behavior of Zn in bone applications to address those contradicting results.

3. Current Orthopedic and Dental Implant Applications of Metallic Wires


The skeletal system of the human body as a complex three-dimensional structure plays two
main roles, supporting the many body organs and their related tissues and attaching the numerous
muscle groups which are needed for body movements. Bone as the hard part of the skeleton is a
natural two- phase organic-inorganic ceramic composite. The organic phase gives bone its flexibility,
while the inorganic phase provides bone with its structural rigidity [72]. The combination of organic
and inorganic phases in the matrix provides bone with the unique mechanical properties such as
toughness, strength, and stiffness. These properties give the ability to withstand against various
mechanical loads applied during the normal and intense physical activities. Despite the remarkable
mechanical and structural characteristics, the bone may fracture due to various loads resulting from
sudden injuries and repeated cyclic loads. An effective solution in restoring the functionality of
damaged bone is surgical implantation of artificial biomaterials. Several types of bone fixation
devices in the form of screws, plates, nails, staples, and wires are produced for orthopedic and
dental implant applications. The selection of biomaterials for these devices is highly dependent on
the specific medical application. The new biomaterials which were introduced and developed for
implant applications should have excellent biocompatibility, comparable strength to natural bone and
produce no cytotoxicity effects [73]. The present review paper focuses on the bone application of wires;
therefore, in this section, bone fixation devices which are made from wires are introduced first and
then the evolution of the biomaterials to be compatible with natural bone are explained.

3.1. K-Wires
Martin Krischner (1879–1942) was a German surgeon who developed a device that allowed him
to insert chromed piano wire from 0.7 to 1.5 mm in diameter into fractured bones to be used as a
traction anchor instead of the large Steinmann pins [74]. Krischner wires (K-wires) were never used by
their inventor for fracture fixation and orthopedic implants [75]. Otto Lowe [74] in 1932 published the
first paper describing the use of K-wires for fracture fragment stabilization. Nowadays, these wires
with different sizes have been used to hold bone fragments together (pin fixation) or to provide an
anchor for skeletal traction. K-wires can also serve as guide pins for the placement of cannulated

115
Metals 2018, 8, 212

screws and can be placed in bone either by hand or at low speed with a power drill [75,76]. Figure 4a
shows radiographic picture of K-wires used to repair a fracture of the medial epicondyle in the elbow.

Figure 4. (a) Radiographic picture of K-wires used to repair a fracture of the medial epicondyle in the
elbow and (b) radiograph shows cerclage wiring used to contain bone fragments in a fracture around
the stem of a femoral prosthesis, reproduced with permission from [76], The Radiological Society of
North America (RSNA® ), 1991.

3.2. Cerclage Wires


The other application of thinner orthopedic wire is as a means of stabilizing fracture fragments.
Cerclage wiring refers to encircling fragments and tensioning the wire to hold fragments in
alignment [76]. A single or double strand of the wire is placed around the bone, and then the
ends are twisted together. Cerclage wires are usually used in combination with other types of fixation
devices such as plates or nails, but can also be used alone in special situations [76]. The radiograph in
Figure 4b shows cerclage wiring used to contain bone fragments in a fracture around the stem of a
femoral prosthesis.

3.3. Tension-Band Wires


The tension band technique converts a tensile force to the compression one. Tension banding is
particularly useful in the setting of fractures where a muscle pull produces distraction of the fracture
fragments, such as fracture of the patella [76]. Parallel K-wires are placed to provide rotational stability
and reduce shearing forces between the fragments. Figure 5 presents an application of tension band
wiring in repairing of an olecranon fracture.

116
Metals 2018, 8, 212

Figure 5. The radiograph shows tension band wiring used to repair an olecranon fracture, reproduced
with permission from [76], The Radiological Society of North America (RSNA® ), 1991.

3.4. Orthodontic Archwires


An orthodontic archwire is a wire conforming the dental arch and exerts forces to correct
irregularities in the position of teeth. An archwire can be used with dental braces as a source of
force. Orthodontic wires release the energy stored upon its placement by applying forces and torque
to the teeth through the appliances placed on them [77]. The forces are applied to move the teeth to a
targeted position. Another application of these wires is maintaining the existing dental position; in
this case, they have a retentive purpose [77].

3.5. Ligature Wires


Ligature wires are small flexible and twistable ties or rings, which are used widely in dentistry
to hold or ligate the archwires to the brackets [9]. This kind of wires is mostly made from SS and
elastomer [78].

3.6. Staples
The bone staple is essentially simple in design and use. This device is made up two or more
points of entry into bone that are attached to each other. They can be fabricated by double bending a
pin or wire that is biased cut to create the sharp ends or tips [10]. The bone staples are fixed into the
bone to help stabilize a fracture to promote bone healing. Bone staples can be used as adjuncts to other
forms of fixations or as single or multiple bone staples at one site.

3.7. Sutures
Metallic sutures are available in many sizes and in two forms: monofilament or multi-strand.
Most surgeons find the multi-strand wires are easier to handle since they are pliable and ties like
silk [79].

4. The evolution of Metallic Wires in Bone Fixation Devices


K-wires as a versatile tool in the hand of orthopedic surgeons have been used for temporary or
definitive osteosynthesis, temporary joint transfixation or for the guidance of other implants such as
cannulated screws. K-wires are present in different forms and designs. Traditional K-wires have a

117
Metals 2018, 8, 212

circular cross-section and two tip designs (diamond and trocar tips). Diamond tip K-wires require the
least axial force to penetrate the bone and consequently generate low heat during drilling, while the
trocar tip wires need a greater force and generate higher temperature on insertion [80]. According to a
surgical instrumentation catalogue, the diameter of standard K-wires based on their applications varies
between 0.4 and 3.5 mm. Besides the strengthening of K-wires by enhancing the mechanical properties
of their materials, the configuration of them has been studied in some researches [81,82]. Different
K-wire fixation techniques contained two or four K-wires with different diameter were investigated
and it was observed that the type of configuration was an effective parameter in increasing of the
rigidity and stability of the fractured bone [81,82].
The first usage of wires in fracture fixation was reported in 1775 when Fe wires were used in a
bone surgery [27]. In 1906, Lambotte used a combination of a Fe wire cerclage, Mg plate and steel
screws to stabilize the fracture of the lower leg [17]. The occurrence of an electrochemical reaction
between Mg and Fe caused extensive subcutaneous gas cavities, local swelling and pain one-day
post-surgery [27]. By observing these symptoms, Lambotte learned that to avoid galvanic corrosion of
the Mg, it should not be implanted with other metals [17]. Lambotte also used pure Mg nails alone to
fix supracondylar fractures of four children. He observed the formation of gas cavities due to the high
corrosion rate of magnesium in vivo [37]. Mainly, because of this problem, the application of Mg wires
was restricted, and SS wires were introduced instead. The biometals used in wire bone fixation devices
are classified into two categories, permanent and biodegradable metals. Among permanent metallic
wires, surgical SS, cobalt-chromium (CoCr) alloys and Ti alloys are the most common ones. In the
following sections, the application of these two groups was described in wire bone fixation devices.

4.1. Permanent Metallic Wires


In 1920, the first type of SS namely 18-8 was used as a bone implant. This alloy showed higher
corrosion resistant and mechanical strength in comparison with that of pure Fe. Since then, various
types of SSs were developed and some of them were used as a proper material for K-wires and cerclage
wires. In 1962, the American Society for Testing and Materials (ASTM) has formed committee F04 on
medical and surgical materials and devices to standardized metals for medical applications. Table 2
shows the differently registered biomaterials for bone wires in the ASTM standard.
In 1988, Stark et al. were the first researchers to recommend the use of SS K-wires in all grafted
fractures, since they were easy to accomplish and required short operating time [83]. They also
achieved 97% union of long-standing ununited fractures of the scaphoid by examining 151 patients
through 22 years, which was a great achievement with K-wires [83]. SS K-wires provide strong fixation
but can be difficult to manipulate and often requires a secondary operation to remove the wire. Due to
the mismatch between the mechanical properties of the wires and the natural bone, mechanical forces
and loads are retained by the wires and are not transferred to the damaged bone. Consequently, stress
shielding happens and results in bone resorption and wire loosening. In addition to stress shielding
effect as well as need for the second surgery, K-wires may cause percutaneous pin tract infection and
wire migration [84,85].

Table 2. Common permanent biomaterials used for bone wires, data from [86].

Chemical Composition Alloy ASTM Number Application


Fe-18Cr-14Ni-2.5Mo (C < 0.03) 316 L F138-08/F139-08/F1350-08 Kirschner wires, cerclage wires
Fe-19Cr-10Ni-2Mn 302 A313/F899 guidewires
Fe-20Cr-10.5Ni-2Mn - A313/F899 Guidewires, orthodontic wires
Fe-20Cr-10.5Ni-2Mn (C < 0.03) - A580 Guidewires, Orthodontic wires
35Co-35Ni-20Cr-10Mo (Ti = 0.7) MP35N F562 Orthopedic wires
35Co-35Ni-10Mo (Ti = 0.01) 35NLT F562 Orthopedic wires
40Co-20Cr-16Fe-15Ni-7Mo Elgiloy F1058–08 Surgical wires
Ti-6Al-4V - F136 Kirschner-wires
Ti-Ni Nitinol F2063 Archwires, Kirschner-wires

118
Metals 2018, 8, 212

The most basic materials used in K-wires are SSs, Co-Cr alloys, Ti and its alloys and nitinol.
It should be noted that the nitinol is an alloy of nickel and titanium elements where the two elements
are present in roughly equal atomic percentages. Since 1930, it has been known that Co-based alloys
due to their excellent wear resistance which is higher than SS, are proper candidates for bone repair
devices [87]. These alloys possess excellent mechanical properties. CoCr alloys also have one order
of magnitude greater corrosion resistance in comparison with SSs. Mainly, due to these properties,
cobalt and its alloys specially, MP35N, modified 35NLT and Elgiloy, have achieved wide attention for
fabrication of K-wires, cerclage wires and guide wires [86].
The performance of medical grade Ti alloys is superior to that of SS and cobalt alloys, due to
the 50% greater strength to weight ratio of the former. These alloys have a lower modulus, higher
corrosion resistance and better biocompatibility in comparison with SSs and cobalt-based alloys. These
advantages make the Ti and its alloys a better-suited alternative for bone applications with higher
loading rates [88]. As compared to commercial unalloyed Ti, Ti-6Al-4V alloy has greater strength, it
has been selected as a proper material for K-wires and cerclage wires. The application of this alloy has
been approved by the ASTM for wire bone fixation. Therefore, various medical device manufacturers
utilize Ti-6Al-4V alloy for producing K-wires in a wide range of diameters. Clauss et al. have treated
135 toe deformities using both SS K-wires and Ti K-wires for one year. They observed that the Ti
K-wires represent less recurrence of deformity, pain, and biofilm formation than SS k-wires. Hence, the
authors suggested utilization of Ti K-wires instead of SS ones for transfixation of toe deformities [89].
Despite the lower density of Ti alloys, their wear resistance and bending strength are lower than cobalt
alloys [87].
Nitinol has proper mechanical stability, lower stiffness, and thermo-elasticity as well as acceptable
corrosion resistance. These properties make it a suitable substitution for SS implants [88]. Fracture
fixation by SS K-wires provides more bending stiffness in comparison with the nitinol wires, while
nitinol wires are more durable. Additionally, nitinol typically costs about 4 times as much as SS with
all else equal.

4.2. Biodegradable Polymeric (BP) Wires


In order to eliminate the secondary surgery and resolve the shortcomings of permanent metallic
wires, the fracture fixation with biodegradable wires made of polymeric materials was suggested and
applied in the clinical trials [90,91]. The wires made of biodegradable synthetic polymers were applied
in the fracture fixation of the knee [91], wrist [92], elbow [84] olecranon and patella [93] and small bones
of hand [90] and their usage were compared with the metallic K-wires. It was concluded that in both
cases the healing time was the same and polymeric wires were suitable alternatives for the metallic
ones [84,90,91,93]. Despite the high cost of polymeric biomaterials, the total expenses of treatment
by polymeric wires was estimated lower than the metallic ones, since there is no need for the second
surgery [93]. In the fixation of wrist fracture, treatment by BP wires was not recommended, because of
high complication rate in patients [92]. In other cases, it was suggested to use polymeric wires as an
adjunct to metal fixations [94,95]. It has been reported that BP materials have a common disadvantage
in load-bearing applications according to their innate mechanical properties. Considering a proper
mechanical strength, biodegradable metals are promising candidates in load-bearing situations, where
a high mechanical strength and a suitable Young’s modulus are required.

5. Recent Developments of BM Wires in Bone Applications


Although the idea of deployment of BM wires has been presented more than a century ago, their
development has been retarded until nearly a decade ago. In this section, the evolution of BM wires
has been introduced with the only focus on Mg, Fe, and Zn and their alloys.
The first experimentally employed Mg wires were used as sutures to anchor nerves and muscles.
During the early attempts, the lack of ductility prevented the use of Mg wires as suture [19]. Although
the pure Mg as a suture wire presented an adequate degradation rate [19], rapid degradation was an

119
Metals 2018, 8, 212

obstacle to develop the usage of Mg as an implant in bone surgery. In orthopedic applications, implants
usually undertake a certain load during the healing of injured bone. Due to the rapid degradation, the
mechanical strength and integrity of the Mg-based implants will be deteriorated seriously. Hence, the
implants lose their load-bearing ability during their service time. Pits or cracks created by the rapid
corrosion might result in sudden failure, even at the initial stage after implantation [96]. Therefore,
it can be concluded that as a bone fixation wire material, two main drawbacks of Mg could be low
ductility and high degradation rate.
Recent studies on Mg-based BMs have focused on the improvement of mechanical properties
and corrosion resistance of Mg and its alloys. In this regard, various methods such as selection on
alloying elements [97–100], microstructural adjustment [101–104] and surface modification [96,105]
have been applied. All the mentioned methods have been widely investigated for examining the in vivo
and in vitro characteristics of common implant devices such as screws, plates, and pins [106–108].
In addition to the clinical trials of McBride and Verbrugge [109], at the first half of the 20th century, to
employ Mg screws, plates and nails, recently, a few researchers [110,111] developed Mg-based implant
devices and utilized them in patients’0 body to fix the bone fractures. Although plenty of studies
has investigated the application of BP materials for manufacturing K-wires and cerclage wires, less
attention has been paid to develop BMs for these wire bone fixation devices. By analyzing the plenty
of publications about biodegradable Mg-based materials, it has been found that just a few pieces of
research have investigated the development of the Mg alloys in the form of wires for bone fraction
surgeries. As it was mentioned, the diameter of K-wires and cerclage wires vary between 0.4 and
3.5 mm. Hence, in this section, only the wires in this range of diameters were considered.
Various alloying elements were added to pure Mg to reduce the corrosion rate and concurrently
keep the biocompatibility of the alloy for bone surgery. Among these alloys, Mg-calcium (Ca) alloys
were biocompatible but degraded quite rapidly [97]. The combination of Mg with rare earth metals
decreased the corrosion rate but also created cytotoxic effects [97,99]. Tie et al. [112] introduced a novel
Mg alloy containing 2% silver (Mg2Ag) and showed that the alloy possessed appropriate mechanical
properties and a rather low degradation rate in vitro. In order to assess the in vivo behavior of the alloy,
Jahn et al. [113] implanted Mg2Ag intramedullary wires into mice with and without a femoral shaft
fracture. The wires with the final diameter of 0.8 mm were produced via the cold rolling following
the hot extrusion. To avoid the loss of ductility, heat treatment between the mentioned metal forming
processes were employed. As a matter of biological effects, Mg2Ag alloy does not have any side effects
on the growth rate of the bone or on inner organ morphology that represents the high biocompatibility
of this alloy. They have examined the degradation rate of the alloy both in vitro and in vivo. Although
the corrosion rate of Mg2Ag alloy in vivo was three times faster than in vitro, the alloy still showed
sufficient rate of biodegradability which maintained proper mechanical stability while supporting
fracture healing. Figure 6 illustrates the X-ray images of fractured bone fixed by steel and Mg2Ag
wires during 133 days in mice [113]. Finally, they have concluded that Mg2Ag might be a promising
material that has a potential application in musculoskeletal medicine.
In another attempt, Bian et al. introduced Mg-based BMs with dietary trace element germanium
(Ge) for orthopedic implant application [114]. They added various mass percentages of Ge to Mg
but in a limited range because of biosafety concerns. The MgGe samples with a diameter of 2.2 mm
were prepared by hot rolling and machining. The highest strength and elongation were reported for
Mg3Ge alloy with the yield tensile strength (YTS) of 135 MPa, UTS of 236 MPa and 17.7% elongation.
It could be concluded that in comparison to other Mg-X alloy (X represents for essential/possibly
essential element in human health) Mg-3Ge exhibits the highest strength and significantly improves
elongation and serves the lowest in vivo corrosion rate. Figure 7 gives a comparative diagram to assess
the strength and elongation of various biodegradable Mg-X alloys manufactured in different process
conditions [114].

120
Metals 2018, 8, 212

Figure 6. X-ray images to determine degradation of intramedullary Mg2Ag wires during 133 days in
mice compared to a steel pin (scale bar = 5 mm), reproduced with permission from [113], Elsevier, 2016.

Figure 7. Comparison of the mechanical properties among various biodegradable Mg-X alloys
(X: essential/possibly essential element in human health). Different symbols represent different
alloy systems and symbol color reflects the process condition, reproduced with permission from [114],
Elsevier, 2017.

121
Metals 2018, 8, 212

Drawing and extrusion are the most common industrial processes for the fabrication of wires
based on the desired diameters. Bian et al. [114] have produced the Mg3Ge wires by machining
following hot rolling process which is not a proper method for manufacturing wires. It is clear
from Figure 7 that extruded parts for most of the Mg alloys have a higher ductility and strength in
comparison with rolled and casted parts. Considering the last two points, one can conclude that more
ductility and higher strength would be achieved by the fabrication of Mg3Ge wires via wire extrusion.
The relevance between the manufacturing process and the corrosion rate for Mg alloys have been
investigated by Koo et al. [115]. They fabricated as-cast Mg-Zn-Mn and as-extruded Mg-Zn-Mn wires
with a diameter of 1.6 mm. Both the alloys were examined in vivo and in vitro. It was observed that
as-cast Mg-Zn-Mn alloy had higher corrosion rate with severely localized pattern in comparison with
the extruded one which had the uniformly localized pattern. This observation has been attributed
to the larger grains boundary and the lower mechanical strength of the as-cast alloy compared to
the extruded one which had smaller grain size and higher mechanical strength [115]. This research
highlighted the effect of the manufacturing process on the mechanical properties and the corrosion rate
of the BM alloys. Therefore, it can be stated that by imposing higher strains during the manufacturing
process of wires, finer microstructure could be achieved, which in turn improves the mechanical
properties and corrosion rate of BMs. Despite the limited studies about Mg wires applied in bone
fixation devices, it could be stated that Mg wires have a high potential for utilization as K-wires and
guide wires in orthopedic surgeries.
In order to overcome the low corrosion rate of pure Fe, various alloying elements such as Mn,
Co, Al, W, Sn, B, C, S and Si have been introduced to accelerate its degradation rate [116–118]. After
adding Mn and Si into pure Fe (Fe30Mn6Si), its corrosion rate was increased [118], while adding
Co, Al, W, Sn, B, C and S have no appreciable influences on the degradation rate of pure Fe [116].
All the mentioned elements can increase the strength of the base metal except Sn which decreases
the mechanical properties [116]. Fe-based alloys also have high strength and high ductility, making
them suitable candidates for utilization as wire fixation devices. Additionally, the superior mechanical
properties of these alloys are favored for making fine wires even in sub-micron sizes which could be
beneficial for the plenty of clinical operations. However, the high Young’s modulus of these alloys
offers an inadequate match to those of natural bone and potentially causes stress shielding.
Hermawan et al. [119] reported the corrosion rate of Fe-Mn alloy which was finally one order of
magnitude lower than that of Mg and its alloys. To achieve high degradation rates, Fe-Mn-Pd alloys
were produced and showed higher corrosion rate compared to pure Fe and Fe-Mn alloys [51]. It is
worth mentioning that these researchers obtained the corrosion rate of those alloys in vitro, which
in fact are different from in vivo results. Kraus et al. [48] examined Fe-Mn-Pd alloy wires in vivo to
perform a feasibility study of these alloys in osteosynthesis applications. They produced the wires
with a diameter of 1.6 mm by casting and machining processes. The degradation rate of Fe-Mn-Pd
wires was not significantly higher than that of Fe and Fe-Mn alloys (see Figure 8).
Attempting to find a proper Fe-based alloy with suitable degradation rate, some researchers
achieved novel Fe alloys such as Fe-Pd [120], Fe-Mn-C-S [121] and Fe-Ga [122] which showed higher
corrosion rate compared to pure Fe in vitro. Capek et al. [120] stated that the preparation route
significantly influenced material properties such as mechanical strength and corrosion rate. In spite of
numerous studies for improving the degradation rate of Fe-based alloys have been reported, only one
research in the literature [48] has applied Fe-based wires in vivo experiments in the rats’ long bone.
As discussed in Section 2.3, Zn and its alloys have moderate corrosion rate and biologically
beneficial corrosion products thus have been considered as one of the first priorities in the material
selection for the regeneration of bone defects. However, there is no in vivo studies in the literature
about the utilization of Zn-based wires in the orthopedic or dental applications. Actually, the
contribution of Zn-based wires in vivo experiments has been mostly limited to the cardiovascular
stent application which has different biological environment compared to bone tissues. Anyway, the
alloying elements, process conditions for wire drawing, mechanical properties, and the corrosion

122
Metals 2018, 8, 212

behavior of Zn-based stents could be an initial guidance for the future use of this kind wires in the
bone regeneration purposes.

Figure 8. Optical micrographs of the wires after implantation for (a–c) 4 weeks and (d–f) 52 weeks of
(a,d) pure Fe, (b,e) Fe–10Mn–1Pd and (c,f) Fe–21Mn–0.7C–1Pd, reproduced with permission from [48],
Elsevier, 2014.

5.1. Medical Sutures and Staples


In 1849, Sims successfully used a fine structure of drawn silver wires by a jeweler to close a
large vesicovaginal fistula [123]. In 1947 Babcock et al. [123] published their experimental results for
utilization of tantalum (Ta) and SS wires as sutures. They found that both the materials were suitable
candidates for suturing but the 18-8 SS which had a reasonable price could be more applicable [123].
For many years and until now, SS sutures have gained a lot of attentions. Permanent metals adopted
for utilization as sutures were mentioned in Table 2. Since the beginning of the 2nd world war, BP
materials were introduced. The BP materials could be classified in two groups, synthetic polymers
(nylon, polyester and propylene) and natural polymers (catgut and collagen) [124]. Materials that were
introduced for suture applications should have high mechanical strength and integrity and should be
biocompatible. It has been reported that polymers cannot withstand high loads securely and exhibit a
low creep and stress relaxation resistance which limits their range of application [13]. Metallic sutures
possess, in general, the higher strength and more advantageous creep and relaxation behavior than
polymers, and they also show, in most of cases, good biocompatibility. Therefore, metallic sutures are
preferred to polymeric ones especially when a biodegradable metallic suture is applied. Mg, Fe and
Zn as biodegradable metals could attract more attention for suture applications.
The first usage of Mg as a suture material dates to the beginning of the 20th century. There were
some attempts to use Mg wires for osteosynthesis and in vascular surgeries. The low ductility of
pure Mg due to its close-packed hexagonal structure has prevented their usage as suture materials.
In 1924, Seeling produced high purity Mg in a distillation process to overcome the low ductility of
the material [13]. The high purity Mg did not exhibit adequate mechanical properties to allow a
defect-free kinking and knotting of the wires. Seeling also tried to increase the ductility of Mg by
adding some alloying elements [13]. The corrosion rate of Mg as suturing material was reported to
be adequate. Hence, Seeling focused on the improvement of its mechanical properties. As it was
mentioned earlier, Mg and its alloys have high corrosion rate which is inappropriate for most of the
applications. Therefore, decreasing the degradation rate of Mg by some methods such as adding
alloying elements were carried out. Researches in this field were retarded until the beginning of
21th century.
Seitz et al. [125,126] were the first researchers produced Mg wires for suture applications. They
selected four Mg alloys, ZEK100, MgCa0.8, AL36 and AX30, which were previously reported to be

123
Metals 2018, 8, 212

biocompatible. Mg suture wires of 0.3 to 0.5 mm in diameter were fabricated by hot extrusion [125]
and their diameter was further decreased via multiple drawing [126]. Despite low mechanical
properties of the extruded wires, they could meet the parameters required for a surgical suture. As an
extruded suture, the AL36 wire had the greatest ductility which allowed for tightening knots (Figure 9).
With the aid of multiple wires drawing process, the Mg wires were reduced to monofilament wire
possessing diameters between 0.5 and 0.1 mm and they also twisted into polyfilament sutures using
stranding. Compared to monofilament strands, polyfilament alloy materials possessed significantly
lower tensile strengths but higher fracture strains [126]. More ductility along with the adequate
strength of polyfilament alloy materials made them suitable candidates for suture applications.

Figure 9. Knotted noose made of the alloy AL36, reproduced with permission from [125], WILEY, 2010.

The previously mentioned studies have investigated pure Mg in combination with aluminum
(Al) as the alloying element. It has been shown that it could be detrimental to the body. In high
doses, Al has been shown to cause neurotoxicity [26]. Considering the neurotoxicity of element
Al, Mg-Al-based alloys are not recommended candidates for biodegradable implant material [27].
However, recently, Al-free Mg alloys with better biological performance such as Mg-RE (rare earth
element) alloys have been developed and investigated in vivo and in vitro [98,99,127]. Bai et al. [127]
employed hot extrusion and cold drawing with intermediate annealing process to prepare fine wires
with final diameters less than 0.4 mm of three kinds of ternary Mg-4%RE (Gd/Y/Nd)-0.4%Zn alloys.
High yield strength along with moderate and adequate ductility was reported for all the fine wires. The
in vitro degradation performance of the finished fine wires was evaluated, and the results indicated
that Mg-4Gd-0.4Zn and Mg-4Nd-0.4Zn wires showed a similar good corrosion resistance and a
uniform corrosion behavior in simulated body fluid (SBF) solution. In contrast, Mg-4Y-0.4Zn fine
wires exhibited a relatively high degradation rate and pitting corrosion behavior [127]. All the studies
dedicated to wires with the diameter of less than 1 mm have been reviewed in this section. Generally,
one can conclude that Mg-based fine wires might be good candidates for suture applications with
adequate mechanical properties and adequate degradation rate. Another application of fine wires is as
staples. Staples made of pure Mg [128,129] and its alloys [130] were recently investigated for gastric
anastomosis. Although the examined staples have exhibited enough closure strength, homogeneous
corrosion behavior, and desirable biodegradation rate both in vivo and in vitro [128–130], their usage
as a bone staple were not evaluated in any researches yet.
In the 18th century, Icart successfully employed Fe wires for the first time to support the healing
of fractured human bone [13]. At that time, Fe was not considered to be a degradable implant material
and was selected only due to its sufficient mechanical properties. It was just at the beginning of
the 21st century that Fe attracted researchers’ attention as a biodegradable and more importantly
nutrient implant material in bone surgeries. However, the use of pure Fe and Fe-based alloys as a
nutrient suture material were not reported in the literature. Table 3 summarizes all the applications of
biodegradable metallic wires.

124
Table 3. Biodegradable metallic wires in bone applications.

Wire Diameter Type of Biological Published


Base Metal Alloy Fabrication Process Application
(mm) Experiments Time
ZEK100 [125] 0.3–0.5 Hot extrusion Suture - 2010
MgCa0.8 [125] 0.3–0.5 Hot extrusion Suture - 2010
Metals 2018, 8, 212

AL36 [125] 0.3–0.5 Hot extrusion Suture - 2010


AX30 [125] 0.3–0.5 Hot extrusion Suture - 2010
ZEK100 [126] 0.274 m –0.822 p Hot extrusion + Cold drawing + Intermediate annealing Suture - 2011
MgCa0.8 [126] 0.274 m –0.822 p Hot extrusion + Cold drawing + Intermediate annealing Suture - 2011
AL36 [126] 0.274 m –0.822 p Hot extrusion + Cold drawing + Intermediate annealing Suture - 2011
Mg
AX30 [126] 0.274 m –0.822 p Hot extrusion + Cold drawing + Intermediate annealing Suture - 2011
Mg-4Gd-0.4Zn [127] 0.3 Hot extrusion + Cold drawing + Intermediate annealing Suture in vitro 2014
Mg-4Y-0.4Zn [127] 0.3 Hot extrusion + Cold drawing + Intermediate annealing Suture in vitro 2014
Mg-4Nd-0.4Zn [127] 0.3 Hot extrusion + Cold drawing + Intermediate annealing Suture in vitro 2014
Mg2Ag [113] 0.8 Hot extrusion + Cold drawing + Intermediate annealing Guide wire-thin pin in vitro-in vivo (mouse) 2016
Mg3Ge [114] 2.2 Hot rolling + Machining Guide wire-thin pin in vitro-in vivo (rabbit) 2017
Mg-Zn-Mn [115] 1.6 Hot extrusion Guide wire-thin pin in vitro-in vivo (mouse) 2017
Fe–10Mn–1Pd [48] 1.6 Casting + Machining Guide wire-thin pin in vitro-in vivo (rat) 2014
Fe
Fe–21Mn–0.7C–1Pd [48] 1.6 Casting + Machining Guide wire-thin pin in vitro-in vivo (rat) 2014
m: Monofilament, p : polyfilament.

125
Metals 2018, 8, 212

5.2. Bone Tissue Engineering


Bone grafting is known as an operation aiming to replace diseased or injured bones with grafting
materials. The bone graft could be taken from the patient (autograft) or donors (allograft) or utilized
synthetic materials [131]. Even though both of autograft and allograft have some significant advantages,
owing to some clinical problems such as the necessity of performing additional surgery on the donor
site as well as size and geometry limitations of autograft and inconsistency with host tissue, and the
possibility of disease transmission from donated allograft, a growing demand for synthetic materials
as bone substitutes is received from clinicians [132]. Rely on the diamond concept for the bone healing
process, synthetic bone materials could play an important role as a scaffold because it can keep
mechanical stability. In addition, a combination of these materials with osteogenic cells and growth
factors could accelerate the new bone formation [132,133].
Among the wide range of synthetic materials, calcium phosphate cement (CPC) is known to be
the most resemble synthetic material to natural bone thus is clinically suitable for the repair of bone
defects in a variety of orthopedic and dental applications [134,135]. Although CPCs are biodegradable,
biocompatible, osteoconductive and have proper compression strength, because of their low bending
and tensile strength and low fracture toughness their applications are limited to non-load bearing areas
such as small cranial and maxillo-facial surgeries [136,137]. Generally, the mechanical strength of CPC
is similar to that of trabecular bone, or one-fifth of that of cortical bone [138]. Reinforcement of CPC
with fibers is one of the most promising approaches for its application in thin and large bone defects as
well as stress-bearing locations [135]. The fiber reinforced calcium phosphate cement (FRCPC) concept
was firstly introduced by Gonten et al. in 2000 [139]. They utilized a fiber-mesh made from poly
(galectin) to reinforce CPC. Their results showed that the knitted fiber-mesh could enhance the flexural
and toughness of CPCs. Then, in 2002, Xu et al. [140] incorporated two types of resorbable fibers into
CPC with a fiber volume fraction of 25%. Their FRCPC increased the flexural strength threefold, and
work-of-fracture toughness nearly 100 times than those of unreinforced CPC. They also claimed that
resorption of the fiber during the regeneration process could facilitate vacuolization and consequently
rapid bone-ingrowth. In a same period of time, Dos Santos et al. [138] followed up this field using
non-degradable fibers including carbon, nylon which were not suitable for clinical applications. They
just tried to enhance the mechanical properties which CPC is weak at.
In the next generation of FRCPC, biocompatible and biodegradable polymeric fibers such as poly
(lactic-co-glycolic) acid (PLGA) [141] and biopolymers such as chitosan and gelatin [142] were used.
Owing to the lower Young’s modulus and strength of these fibers than those of the cement matrix,
they are not successful enough as a reinforcement element for the strengthening of CPC. The bending
strength of 40–45 MPa is the best mechanical properties have been reported for FRCPC achieved by
PLGA fibers up to now. Furthermore, in physiological environment, fast degradation rate of PLGA
fibers results in the concern of mechanical stability only after 3–4 weeks [143]. In 2013, the new
generation of FRCPC was introduced by Kruger et al. [144]. They utilized AZ31 and ZEK100 Mg wires
with 0.6 and 0.228 mm in diameter respectively. It was for the first time that BM wires were used as a
reinforcement element for the strengthening of CPC which was successful to improve the bending
strength of FRCPC up to 139 ± 41 MPa. They stated that regarding the fracture of the FRCPC initiated
by shearing of the matrix at or near the biodegradable metallic wire surface, hybrid reinforcement of
CPC with both metal and fine polymer fibers such as polylactic acid (PLA) could postpone the crack
initiation and further enhance the mechanical properties of FRCPC (Figure 10). Additionally, in vitro
tests with osteoblasts validated cytocompatibility of wire-reinforced CPC composites with BM.

126
Metals 2018, 8, 212

Figure 10. Comparison between three different compositions of FRCPC including AZ31, PLA and
hybrid (both AZ31 and PLA). (a) Load-displacement curves; (b) Bending strength and (c) Fracture
images of the composites, reproduced with permission from [114], Elsevier, 2013.

However, there is only one pilot study for the use of FRCPC composites. Further investigations
need to be done to optimize the mechanical properties of the BM wire reinforced CPC composites by
applying a vast range of diameters, plenty types of biodegradable alloys and different fiber dispositions.
Moreover, assessment of the mechanical stability should be conducted in much longer time in vitro
(several months or even a year), and finally in vivo animal tests should be performed to examine
all of the positive and negative biological effects and structural changing during the degradation of
these kind of composites before any clinical applications for the grafting of human bone defects in the
load-bearing locations.

127
Metals 2018, 8, 212

6. Future Work
Generally, for BM wires and especially for Mg and its alloys, controlling the degradation rate
for implant applications is quite important. It becomes more critical for the application of fine BM
wires in stress-bearing areas, because the mechanical properties of such wires can be significantly
influenced by high corrosion rate. Therefore, the first priority in the application of BM wires is
controlling the corrosion rate to a logical rate, at least in the initial stage after implantation. In this
regard, methods such as selection of alloying elements, microstructure adjustments and surface
modification will be effective. It is worth mentioning that these corrosion controlling methods have
some intrinsic differences. Various surface modification techniques create a protective layer on the
material which resist against corrosion. In some cases, this layer also increases the osteoinductivity
and osteoconductivity of the material while in other cases only controls the corrosion rate on the
surface. In general, surface modification methods affect the surface corrosion. Selecting proper alloying
elements in combination with the base metal could adjust the volume degradation rate of the material
and concurrently could enhance mechanical properties of an alloy. The third corrosion controlling
method, by employing a proper fabrication process, creates a fine and uniform microstructure which
in turn influences the mechanical properties of the material. Hence, it is obvious that selecting alloying
elements and refining microstructure methods will have another beneficial effect on mechanical
characteristics of the materials. These two mentioned techniques, except in some special cases, are
preferred to surface modification methods.

6.1. Alloying Elements

6.1.1. Magnesium
Common alloying elements used for pure Mg are mentioned in Table 4. Biological and mechanical
advantages and disadvantages of the elements are summarized in this table. Since Al and Cu may
cause neurotoxicity in the body, their usage in recent studies has been limited. It is obvious from the
table that Cu also accelerates the corrosion rate of Mg [26] and hence, it is not suitable for implants.
Adding Ca, Si and Li elements could deteriorate the degradation rate of Mg [97]. Therefore, these
elements cannot play their role as a corrosion controlling element. It can be seen that other alloying
elements, such as Zn, Zr, Mn and RE improve the corrosion resistance of Mg and simultaneously
increase the strength and ductility of the obtained alloy. Additionally, for the mentioned elements in
standard dosage, harmful side effects were not reported [97]. In recent literature it can be found that
newly introduced biocompatible Mg alloys mostly contain these elements [145–148]. Therefore, in
designing and developing of new Mg biomaterials the potentials of these alloying elements should
be considered.

6.1.2. Iron
Alloying elements such as Mn, Co, Al, W, Sn, B, C, S and Si were added into Fe-based alloys to
accelerate their degradation rate. It has been found that adding Co, Al, W, Sn, B, C and S have no
influence on the degradation behavior of pure Fe [116], while the alloy Fe30Mn6Si had a relatively
more corrosion rate compared to pure Fe [118]. There are two main criteria regarding increase in the
corrosion rate of Fe-based alloys which affect the corrosion susceptibility of this kind of alloys: (1) the
addition of less noble alloying elements within the solubility limit in Fe to cause the matrix more
susceptible to corrosion; and (2) the addition of noble alloying elements to generate small and finely
dispersed intermetallic phases (IMPs). Mn and Pd have been shown to be suitable alloying elements
according to both of the mentioned approaches [51]. Fe-Mn-Pd alloy has investigated and suggested as
a Fe-based alloy which could improve the degradation rate of the base metal [48]. Inspecting the effect
of alloying elements on the corrosion behavior of Fe-based alloys is still a challenging and ongoing
research area.

128
Table 4. Biological and mechanical advantages and disadvantages of alloying elements for Mg.

Alloying Element Advantages Disadvantages


Increase the strength and ductility and decrease the
Al Neurotoxicity and accumulation in bone [87].
corrosion rate [26].
Metals 2018, 8, 212

Most abundant mineral and mainly stored in bone and teeth;


Ca Calcium metabolism disorder; kidney stones; increasing corrosion rate [97].
activator or stabilizer of enzymes [87].
Neurodegenerative diseases including Alzheimer’s, Menkes, and Wilson
Cu Increase the ductility [97].
disease; accelerate corrosion and decrease the strength [26].
Essential trace element; appear in all enzyme classes;
Zn Euro toxic and hinder bone development at higher concentration [87].
increase the strength and improve corrosion resistance [87].
Improve corrosion resistance, strength and ductility;
essential trace element; activator of enzyme; Mn deficiency
Mn Excessive Mn results in neurotoxicity [87].
is related to osteoporosis, diabetes mellitus, and
atherosclerosis [87].
Cross linking agent of connective tissue basement Excessive SiO2 causes lung diseases [87]; decreasing the ductility and
Si
membrane structures; necessary for bone calcification [87]. increasing corrosion rate [97].

129
Li Used in the treatment of manic-depressive psychoses [87]. Decrease the strength and increasing corrosion rate [97].
Zr Increase the strength and ductility [97]. High concentration in liver and gall bladder [87].
Compound of drugs for treatment of cancer; improve the
RE Accumulation in bone and liver [87].
mechanical properties and corrosion resistance [87].
Metals 2018, 8, 212

6.1.3. Zinc
Zinc as a newly introduced biodegradable metal with a moderate degradation rate has attracted
some attention. Alloying elements of Zn have the role of strengthening the matrix since Zn-based
alloys suffer from low mechanical properties. To improve the mechanical properties of pure Zn,
some elements such as Mg, Ca, Sr, Sn and Fe were added into Zn-based alloys [56,70,117,149].
Vojtech et al. [70] developed as cast Zn-Mg alloys and studied their mechanical properties and corrosion
behavior. Their results showed a significantly improvement in the tensile strength of Zn modifying
by 1% Mg. The elongation of Zn-1% Mg alloy was much greater than that of pure Zn but was
only 2%, which is quite low for medical usage [70]. Combination of Mg as an alloying element with
Zn-based alloys especially about 1% is newly introduced and investigated by several researchers [54,55].
Shen et al. [150] produced Zn-1.2% Mg alloy and achieved the high ultimate strength of 362 MPa and
elongation of 21%, which is the highest ductility obtained for biodegradable Zn alloys. The combination
of Zn with other alloying elements is still a challenging topic for researchers.

6.2. Surface Modification

6.2.1. Magnesium
Numerous studies have proved that the functional and mechanical behaviors as well as
chemical and physical properties of biomaterials can be affected by different surface modification
techniques [151]. Although plenty of coating methods such as electrodeposition [152], sol-gel and
dipping [153], chemical conversion [154], anodization [155], vapor deposition [156], spin [157], and
alkali treatment [158] have been proposed to inhibit rapid corrosion of Mg and its alloys, just few of
them are effective and flexible enough for the formation of a uniform coating with desired thickness
on the non-flat surfaces such as wires. Simultaneously, the adhesion between the substrate and the
coating film is another concern that should be considered. Finally, prevention from deteriorating the
mechanical properties of the substrate as much as possible is the third consideration in the selection of
an adequate coating method. To summarize, it is claimed that only sol-gel preparation combined with
dip coating and alkali treatment can meet all of the three mentioned requirements which make them
suitable and cost-effective methods for the coating of Mg-based wires [151,159].
The composition of the coating is another critically important parameter. An ideal coating material
for bone applications should be able to enhance the biocompatibility and osteointegration properties
of Mg-based alloys and simultaneously increase the corrosion resistance. As discussed in Section 2.1.1,
coating with Mg(OH)2 [159,160] could enhance bone formation and passivate the Mg-based wires
to decrease the corrosion rate as well. The synthetic calcium phosphate coatings [151,161,162] can
definitely facilitate the new bone formation. The biomimetic coating with pure collagen [163] and
collagen-based composites [164] is another promising approach that can be applied in the future
studies of Mg-based wires in bone tissue regeneration.

6.2.2. Iron
Despite numerous investigations focused on the surface treatment methods for Mg-based alloys,
only few research can be found for surface treatment of pure Fe. Chen et al. [165] suggested using
micro-patterned Au disc arrays to accelerate degradation of pure Fe. The proposed approach
relied on the galvanic corrosion caused by the Au array [165]. In another research, Fe-O thin
films were prepared on pure Fe by plasma immersion ion implantation and deposition in order
to improve the biodegradability and biocompatibility [166]. The prepared thin films decreased the
degradation rate and enhanced the corrosion behavior [166]. Sandblasting method has been proposed
by Zhou et al. [167] as a successful surface treatment for increasing the degradation rate of pure Fe in
SBF. The reasons for increasing the corrosion rate were the change of surface composition (in the early
stage), high roughness and high density of dislocations [167]. The limited number of researches in the

130
Metals 2018, 8, 212

surface treatment methods for Fe-based BM wires indicates that more studies is required to introduce
and reveal the potentials of various surface treatments approaches.

6.3. Fabrication Processes


As stated before, microstructure adjustment is one of the corrosion control methods. In this
regard, selection of a proper fabrication process could effectively influence the microstructure. In case
of Mg-Zn-Mn alloy, it has been reported that the extruded alloy compared to casted material had
fine grain size and accordingly slower degradation rate. Therefore, selection of the manufacturing
process of the alloy needs more attention. There are several common manufacturing processes such
as rolling, extrusion, forging, drawing, and continuous casting. Recently severe plastic deformation
(SPD) processes such as equal channel angular extrusion (ECAP), multi-axial forging (MAF) and
high-pressure torsion (HPT) were introduced. These SPD methods by imposing large shear strains
or huge compressive strains decrease the grain size of the materials. Dobatkin et al. [168] employed
three different SPD processes (ECAP, MAF, and rotary swaging (RS)) to investigate the effects on the
microstructure and corrosion behavior of WE43 Mg alloy. They observed that all the processes could
decrease the grain size and create ultra-fine grained (UFG) microstructure. The biodegradation rate
of the UFG WE43 was examined in vitro and it was concluded that ECAP and MAF caused some
deceleration of biodegradation rate by slowing down the gas formation in the biological fluid and
improved the biocompatibility of the WE43 alloy [168]. This research shows the potential of SPD
method in controlling the corrosion of Mg alloys which can be spread to other biodegradable metals
as well.

7. Conclusions
Application of biodegradable metals (magnesium, iron, and zinc) as promising biodegradable
materials is spreading in various biomedical devices especially for dental and orthopedic applications.
Wire is one of the most popular forms of these materials that can be easily formed by drawing
and other methods. Although currently biodegradable metallic wires have some drawbacks such
as unsatisfactory degradation rate and mechanical properties when used in bone surgeries, many
methods can be adopted to control their corrosion behavior and tailor their mechanical properties.
Owing to the proper mechanical properties, controllable degradation rate, and the release of nutrient
ions, which will significantly be beneficial to bone cells and bone tissue regeneration, we firmly believe
that in the near future BM wires will be used broadly in dental and orthopedic fields.

Author Contributions: Mohammad Asgari, Zhiyong Li and Yin Xiao conceptualized the paper and gave
instructions; Mohammad Asgari wrote the paper; Ruiqiang Hang, Zhiyong Li and Yin Xiao edited the paper;
Mohammad Asgari, Zhiyong Li and Yin Xiao revised the paper; Chang Wang and Zhentao Yu gave idea and
advice; Zhiyong Li and Yin Xiao gave directions.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Griebel, A.J.; Schaffer, J.E. Absorbable filament technologies: Wire-drawing to enable next-generation medical
devices. In Magnesium Technology 2016; Springer: Berlin, Germany, 2016; pp. 323–327.
2. Sardeshmukh, A.; Reddy, S.; Gautham, B.; Joshi, A.; Panchal, J. A data science approach for analysis of
multi-pass wire drawing. In Proceedings of the ASME 2017 International Design Engineering Technical
Conferences and Computers and Information in Engineering Conference, Cleveland, OH, USA, 6–9 August
2017; American Society of Mechanical Engineers: New York, NY, USA, 2017; p. V001T002A071. [CrossRef]
3. Griebel, A.J.; Schaffer, J.E.; Hopkins, T.M.; Alghalayini, A.; Mkorombindo, T.; Ojo, K.O.; Xu, Z.;
Little, K.J.; Pixley, S.K. An in vitro and in vivo characterization of fine WE43B magnesium wire with varied
thermomechanical processing conditions. J. Biomed. Mater. Res. Part B Appl. Biomater. 2017. [CrossRef]
[PubMed]

131
Metals 2018, 8, 212

4. Duerig, T.; Pelton, A.; Stöckel, D. An overview of nitinol medical applications. Mater. Sci. Eng. A 1999, 273,
149–160. [CrossRef]
5. Welch, J.M.; Sutton, G. Coronary Guide Catheter. Google Patents WO2017020012 A1, 29 July 2016.
6. Henkes, H.; Bose, A.; Felber, S.; Miloslavski, E.; Berg-Dammer, E.; Kühne, D. Endovascular coil occlusion of
intracranial aneurysms assisted by a novel self-expandable nitinol microstent (neuroform). Int. Neuroradiol.
2002, 8, 107–119. [CrossRef] [PubMed]
7. Tan, L.; Sun, D.-H.; Yu, T.; Wang, L.; Zhu, D.; Li, Y.-H. Death due to intra-aortic migration of kirschner wire
from the clavicle: A case report and review of the literature. Medicine 2016, 95, e3741. [CrossRef] [PubMed]
8. Bravo, L.A.; de Cabañes, A.G.; Manero, J.M.; Rúperez, E.; Gil, F.J. NiTi superelastic orthodontic archwires
with polyamide coating. J. Mater. Sci. Mater. Med. 2014, 25, 555–560. [CrossRef] [PubMed]
9. Chakravorty, B. Configured ligature wire for quick conventional and overtie lingual archwire ligations.
APOS Trends Orthod. 2017, 7, 108–110. [CrossRef]
10. Iavazzo, C.; Gkegkes, I.D.; Vouloumanou, E.K.; Mamais, I.; Peppas, G.; Falagas, M.E. Sutures versus staples
for the management of surgical wounds: A meta-analysis of randomized controlled trials. Am. Surg. 2011,
77, 1206–1221. [PubMed]
11. Bowen, P.K.; Shearier, E.R.; Zhao, S.; Guillory, R.J.; Zhao, F.; Goldman, J.; Drelich, J.W. Biodegradable metals
for cardiovascular stents: From clinical concerns to recent Zn-alloys. Adv. Healthc. Mater. 2016, 5, 1121–1140.
[CrossRef] [PubMed]
12. Li, H.; Zheng, Y.; Qin, L. Progress of biodegradable metals. Prog. Natl. Sci. Mater. Int. 2014, 24, 414–422.
[CrossRef]
13. Seitz, J.-M.; Durisin, M.; Goldman, J.; Drelich, J.W. Recent advances in biodegradable metals for medical
sutures: A critical review. Adv. Healthc. Mater. 2015, 4, 1915–1936. [CrossRef] [PubMed]
14. Wang, C.; Yang, H.T.; Li, X.; Zheng, Y.F. In vitro evaluation of the feasibility of commercial Zn alloys as
biodegradable metals. J. Mater. Sci. Technol. 2016, 32, 909–918. [CrossRef]
15. Chen, Y.; Zhang, W.; Maitz, M.F.; Chen, M.; Zhang, H.; Mao, J.; Zhao, Y.; Huang, N.; Wan, G. Comparative
corrosion behavior of Zn with Fe and Mg in the course of immersion degradation in phosphate buffered
saline. Corros. Sci. 2016, 111, 541–555. [CrossRef]
16. Li, H.; Yang, H.; Zheng, Y.; Zhou, F.; Qiu, K.; Wang, X. Design and characterizations of novel biodegradable
ternary Zn-based alloys with IIA nutrient alloying elements Mg, Ca and Sr. Mater. Des. 2015, 83, 95–102.
[CrossRef]
17. Witte, F. Reprint of: The history of biodegradable magnesium implants: A review. Acta Biomater. 2015, 23,
S28–S40. [CrossRef] [PubMed]
18. Andrews, E.W. Absorbable metal clips as substitutes for ligatures and deep sutures in wound closure. J. Am.
Med. Assoc. 1917, 69, 278–281. [CrossRef]
19. Seelig, M. A study of magnesium wire as an absorbable suture and ligature material. Arch. Surg. 1924, 8,
669–680. [CrossRef]
20. Hou, L.-D.; Li, Z.; Pan, Y.; Sabir, M.; Zheng, Y.-F.; Li, L. A review on biodegradable materials for cardiovascular
stent application. Front. Mater. Sci. 2016, 10, 238–259. [CrossRef]
21. Li, X.; Chu, C.; Zhou, L.; Bai, J.; Guo, C.; Xue, F.; Lin, P.; Chu, P.K. Fully degradable PLA-based composite
reinforced with 2D-braided Mg wires for orthopedic implants. Compos. Sci. Technol. 2017, 142, 180–188.
[CrossRef]
22. Staiger, M.P.; Pietak, A.M.; Huadmai, J.; Dias, G. Magnesium and its alloys as orthopedic biomaterials:
A review. Biomaterials 2006, 27, 1728–1734. [CrossRef] [PubMed]
23. Zhao, N.; Zhu, D. Application of Mg-based alloys for cardiovascular stents. Int. J. Biomed. Eng. Technol. 2013,
12, 382–398. [CrossRef]
24. Liu, A.; Sun, M.; Shao, H.; Yang, X.; Ma, C.; He, D.; Gao, Q.; Liu, Y.; Yan, S.; Xu, S. The outstanding mechanical
response and bone regeneration capacity of robocast dilute magnesium-doped wollastonite scaffolds in
critical size bone defects. J. Mater. Chem. B 2016, 4, 3945–3958. [CrossRef]
25. Wu, L.; Feyerabend, F.; Schilling, A.F.; Willumeit-Römer, R.; Luthringer, B.J. Effects of extracellular
magnesium extract on the proliferation and differentiation of human osteoblasts and osteoclasts in coculture.
Acta Biomater. 2015, 27, 294–304. [CrossRef] [PubMed]
26. Persaud-Sharma, D.; McGoron, A. Biodegradable magnesium alloys: A review of material development and
applications. J. Biomim. Biomater. Tissue Eng. 2012, 12, 25–39. [CrossRef] [PubMed]

132
Metals 2018, 8, 212

27. Zheng, Y.F.; Gu, X.N.; Witte, F. Biodegradable metals. Mater. Sci. Eng. R Rep. 2014, 77, 1–34. [CrossRef]
28. Pilarska, A.A.; Klapiszewski, Ł.; Jesionowski, T. Recent development in the synthesis, modification and
application of Mg(OH)2 and MgO: A review. Powder Technol. 2017, 319, 373–407. [CrossRef]
29. Guo, M.; Muhammad, F.; Wang, A.; Qi, W.; Wang, N.; Guo, Y.; Wei, Y.; Zhu, G. Magnesium hydroxide
nanoplates: A pH-responsive platform for hydrophobic anticancer drug delivery. J. Mater. Chem. B 2013, 1,
5273–5278. [CrossRef]
30. Janning, C.; Willbold, E.; Vogt, C.; Nellesen, J.; Meyer-Lindenberg, A.; Windhagen, H.; Thorey, F.; Witte, F.
Magnesium hydroxide temporarily enhancing osteoblast activity and decreasing the osteoclast number in
peri-implant bone remodelling. Acta Biomater. 2010, 6, 1861–1868. [CrossRef] [PubMed]
31. Weizbauer, A.; Kieke, M.; Rahim, M.I.; Angrisani, G.L.; Willbold, E.; Diekmann, J.; Flörkemeier, T.;
Windhagen, H.; Müller, P.P.; Behrens, P.; et al. Magnesium-containing layered double hydroxides as
orthopaedic implant coating materials—An in vitro and in vivo study. J. Biomed. Mater. Res. Part B
Appl. Biomater. 2016, 104, 525–531. [CrossRef] [PubMed]
32. Hickey, D.J.; Ercan, B.; Chung, S.; Webster, T.J.; Sun, L.; Geilich, B. Mgo nanocomposites as new antibacterial
materials for orthopedic tissue engineering applications. In Proceedings of the 2014 40th Annual Northeast
Bioengineering Conference (NEBEC), Boston, MA, USA, 25–27 April 2014; pp. 1–2.
33. Huang, L.; Li, D.-Q.; Lin, Y.-J.; Wei, M.; Evans, D.G.; Duan, X. Controllable preparation of Nano-MgO and
investigation of its bactericidal properties. J. Inorg. Biochem. 2005, 99, 986–993. [CrossRef] [PubMed]
34. Haghshenas, M. Mechanical characteristics of biodegradable magnesium matrix composites: A review.
J. Magn. Alloys 2017, 5, 189–201. [CrossRef]
35. Gu, X.-N.; Zheng, Y.-F. A review on magnesium alloys as biodegradable materials. Front. Mater. Sci. China
2010, 4, 111–115. [CrossRef]
36. Yusop, A.; Bakir, A.; Shaharom, N.; Abdul Kadir, M.; Hermawan, H. Porous biodegradable metals for hard
tissue scaffolds: A review. Int. J. Biomater. 2012, 2012, 641430. [CrossRef] [PubMed]
37. Zhao, D.; Witte, F.; Lu, F.; Wang, J.; Li, J.; Qin, L. Current status on clinical applications of magnesium-based
orthopaedic implants: A review from clinical translational perspective. Biomaterials 2017, 112, 287–302.
[CrossRef] [PubMed]
38. Haynes, W.M. CRC Handbook of Chemistry and Physics; CRC Press: Boca Raton, FL, USA, 2014.
39. Feyerabend, F.; Wendel, H.-P.; Mihailova, B.; Heidrich, S.; Agha, N.A.; Bismayer, U.; Willumeit-Römer, R.
Blood compatibility of magnesium and its alloys. Acta Biomater. 2015, 25, 384–394. [CrossRef] [PubMed]
40. Harandi, S.E.; Banerjee, P.C.; Easton, C.D.; Singh Raman, R.K. Influence of bovine serum albumin in hanks’
solution on the corrosion and stress corrosion cracking of a magnesium alloy. Mater. Sci. Eng. C 2017, 80,
335–345. [CrossRef] [PubMed]
41. Noviana, D.; Paramitha, D.; Ulum, M.F.; Hermawan, H. The effect of hydrogen gas evolution of magnesium
implant on the postimplantation mortality of rats. J. Orthop. Transl. 2016, 5, 9–15. [CrossRef]
42. Tang, J.; Wang, J.; Xie, X.; Zhang, P.; Lai, Y.; Li, Y.; Qin, L. Surface coating reduces degradation rate of
magnesium alloy developed for orthopaedic applications. J. Orthop. Transl. 2013, 1, 41–48. [CrossRef]
43. Kraus, T.; Fischerauer, S.F.; Hänzi, A.C.; Uggowitzer, P.J.; Löffler, J.F.; Weinberg, A.M. Magnesium alloys
for temporary implants in osteosynthesis: In vivo studies of their degradation and interaction with bone.
Acta Biomater. 2012, 8, 1230–1238. [CrossRef] [PubMed]
44. Zheng, Y.; Xu, X.; Xu, Z.; Wang, J.; Cai, H. Development of Fe-based degradable metallic biomaterials.
Metallic Biomater. New Direct. Technol. 2017, 113–160.
45. Zimmermann, M.B.; Hurrell, R.F. Nutritional iron deficiency. Lancet 2007, 370, 511–520. [CrossRef]
46. Toxqui, L.; Vaquero, M.P. Chronic iron deficiency as an emerging risk factor for osteoporosis: A hypothesis.
Nutrients 2015, 7, 2324–2344. [CrossRef] [PubMed]
47. Francis, A.; Yang, Y.; Virtanen, S.; Boccaccini, A.R. Iron and Iron-based alloys for temporary cardiovascular
applications. J. Mater. Sci. Mater. Med. 2015, 26, 138. [CrossRef] [PubMed]
48. Kraus, T.; Moszner, F.; Fischerauer, S.; Fiedler, M.; Martinelli, E.; Eichler, J.; Witte, F.; Willbold, E.;
Schinhammer, M.; Meischel, M. Biodegradable Fe-based alloys for use in osteosynthesis: Outcome of
an in vivo study after 52 weeks. Acta Biomater. 2014, 10, 3346–3353. [CrossRef] [PubMed]
49. Huang, T.; Zheng, Y.; Han, Y. Accelerating degradation rate of pure iron by Zinc Ion implantation.
Regener. Biomater. 2016, 3, 205–215. [CrossRef] [PubMed]

133
Metals 2018, 8, 212

50. Mouzou, E.; Paternoster, C.; Tolouei, R.; Chevallier, P.; Biffi, C.A.; Tuissi, A.; Mantovani, D. CO2 -rich
atmosphere strongly affects the degradation of Fe-21Mn-1C for biodegradable metallic implants. Mater. Lett.
2016, 181, 362–366. [CrossRef]
51. Schinhammer, M.; Hanzi, A.C.; Loffler, J.F.; Uggowitzer, P.J. Design strategy for biodegradable Fe-based
alloys for medical applications. Acta Biomater. 2010, 6, 1705–1713. [CrossRef] [PubMed]
52. Keen, C.L.; Zidenberg-Cherr, S. Manganese A2—Caballero, benjamin. In Encyclopedia of Food Sciences and
Nutrition, 2nd ed.; Academic Press: Oxford, UK, 2003; pp. 3686–3691.
53. Dermience, M.; Lognay, G.; Mathieu, F.; Goyens, P. Effects of thirty elements on bone metabolism. J. Trace
Elements Med. Biol. 2015, 32, 86–106. [CrossRef] [PubMed]
54. Sikora-Jasinska, M.; Mostaed, E.; Mostaed, A.; Beanland, R.; Mantovani, D.; Vedani, M. Fabrication,
mechanical properties and in vitro degradation behavior of newly developed Zn Ag alloys for degradable
implant applications. Mater. Sci. Eng. C 2017, 77, 1170–1181. [CrossRef] [PubMed]
55. Katarivas Levy, G.; Goldman, J.; Aghion, E. The prospects of zinc as a structural material for biodegradable
implants—A review paper. Metals 2017, 7, 402. [CrossRef]
56. Li, H.F.; Xie, X.H.; Zheng, Y.F.; Cong, Y.; Zhou, F.Y.; Qiu, K.J.; Wang, X.; Chen, S.H.; Huang, L.; Tian, L.; et al.
Development of biodegradable Zn-1X binary alloys with nutrient alloying elements Mg, Ca and Sr. Sci. Rep.
2015, 5, 10719. [CrossRef] [PubMed]
57. Yang, H.; Wang, C.; Liu, C.; Chen, H.; Wu, Y.; Han, J.; Jia, Z.; Lin, W.; Zhang, D.; Li, W.; et al. Evolution
of the degradation mechanism of pure Zinc stent in the one-year study of rabbit abdominal aorta model.
Biomaterials 2017, 145, 92–105. [CrossRef] [PubMed]
58. Liu, X.; Sun, J.; Yang, Y.; Zhou, F.; Pu, Z.; Li, L.; Zheng, Y. Microstructure, mechanical properties, in vitro
degradation behavior and hemocompatibility of novel Zn–Mg–Sr alloys as biodegradable metals. Mater. Lett.
2016, 162, 242–245. [CrossRef]
59. Zhu, H.; Jin, G.; Cao, H.; Qiao, Y.; Liu, X. Influence of implantation voltage on the biological properties of
Zinc-implanted titanium. Surf. Coat. Technol. 2017, 312, 75–80. [CrossRef]
60. Seo, H.-J.; Cho, Y.-E.; Kim, T.; Shin, H.-I.; Kwun, I.-S. Zinc may increase bone formation through stimulating
cell proliferation, alkaline phosphatase activity and collagen synthesis in osteoblastic MC3T3-E1 cells.
Nutr. Res. Pract. 2010, 4, 356–361. [CrossRef] [PubMed]
61. Moonga, B.S.; Dempster, D.W. Zinc is a potent inhibitor of osteoclastic bone resorption in vitro. J. Bone
Min. Res. 1995, 10, 453–457. [CrossRef] [PubMed]
62. Bhardwaj, P.; Rai, D.V.; Garg, M.L. Zinc inhibits ovariectomy induced microarchitectural changes in the bone
tissue. J. Nutr. Intermed. Metab. 2016, 3, 33–40. [CrossRef]
63. Yamaguchi, M. Role of nutritional zinc in the prevention of osteoporosis. Mol. Cell. Biochem. 2010, 338,
241–254. [CrossRef] [PubMed]
64. Pasquet, J.; Chevalier, Y.; Pelletier, J.; Couval, E.; Bouvier, D.; Bolzinger, M.-A. The contribution of zinc ions to
the antimicrobial activity of zinc oxide. Colloids Surf. A Physicochem. Eng. Asp. 2014, 457, 263–274. [CrossRef]
65. Boyd, D.; Li, H.; Tanner, D.A.; Towler, M.R.; Wall, J.G. The antibacterial effects of zinc ion migration from
Zinc-based glass polyalkenoate cements. J. Mater. Sci. Mater. Med. 2006, 17, 489–494. [CrossRef] [PubMed]
66. McCarthy, T.J.; Zeelie, J.J.; Krause, D.J. The antimicrobial action of Zinc ion/antioxidant combinations. J. Clin.
Pharm. Ther. 1992, 17, 51–54. [CrossRef] [PubMed]
67. Drelich, A.J.; Bowen, P.K.; LaLonde, L.; Goldman, J.; Drelich, J.W. Importance of oxide film in endovascular
biodegradable Zinc stents. Surf. Innov. 2016, 4, 133–140. [CrossRef]
68. Moon, S.-H.; Choi, W.J.; Choi, S.-W.; Kim, E.H.; Kim, J.; Lee, J.-O.; Kim, S.H. Anti-cancer activity of ZnO
chips by sustained Zinc Ion release. Toxicol. Rep. 2016, 3, 430–438. [CrossRef] [PubMed]
69. Choi, S.-W.; Choi, W.J.; Kim, E.H.; Moon, S.-H.; Park, S.-J.; Lee, J.-O.; Kim, S.H. Inflammatory bone resorption
and antiosteosarcoma potentials of Zinc Ion sustained release ZnO chips: Friend or foe? ACS Biomater.
Sci. Eng. 2016, 2, 494–500. [CrossRef]
70. Vojtech, D.; Kubasek, J.; Serak, J.; Novak, P. Mechanical and corrosion properties of newly developed
biodegradable Zn-based alloys for bone fixation. Acta Biomater. 2011, 7, 3515–3522. [CrossRef] [PubMed]
71. Törne, K.; Larsson, M.; Norlin, A.; Weissenrieder, J. Degradation of Zinc in saline solutions, plasma, and
whole blood. J. Biomed. Mater. Res. Part B Appl. Biomater. 2016, 104, 1141–1151. [CrossRef] [PubMed]
72. Katz, E.P.; Li, S.-T. Structure and function of bone collagen fibrils. J. Mol. Biol. 1973, 80, 1–15. [CrossRef]

134
Metals 2018, 8, 212

73. Tan, L.; Yu, X.; Wan, P.; Yang, K. Biodegradable materials for bone repairs: A review. J. Mater. Sci. Technol.
2013, 29, 503–513. [CrossRef]
74. Huber, W. Historical remarks on martin kirschner and the development of the kirschner (K)-wire. Indian J.
Plast. Surg. 2008, 41, 89–92. [CrossRef] [PubMed]
75. Harasen, G. Orthopedic hardware and equipment for the beginner: Part 1. Pins and wires. Can. Vet. J. 2011,
52, 1025–1026. [PubMed]
76. Ricbard, M.; Slone, M.M.H.; Vander Griend Robert, A.; William, J. Montgomery Orthopaedic fixation devices.
RadioGraphics 1991, 11, 823–847.
77. Rudolph, D.J.; Willes, M.G.; Sameshima, G.T. A finite element model of apical force distribution from
orthodontic tooth movement. Angle Orthod. 2001, 71, 127–131. [PubMed]
78. Henriques, J.F.C.; Higa, R.H.; Semenara, N.T.; Janson, G.; Fernandes, T.M.F.; Sathler, R. Evaluation of
deflection forces of orthodontic wires with different ligation types. Braz. Oral Res. 2017, 31, e49. [CrossRef]
[PubMed]
79. Cowley, L.L. Wire sutures: Braided or monofilament? Am. J. Surg. 1967, 113, 472–474. [CrossRef]
80. Karmani, S.; Lam, F. The design and function of surgical drills and k-wires. Curr. Orthop. 2004, 18, 484–490.
[CrossRef]
81. Massengill, J.B.; Alexander, H.; Parson, J.R.; Schecter, M.J. Mechanical analysis of kirschner wire fixation in a
phalangeal model. J. Hand Surg. 1979, 4, 351–356. [CrossRef]
82. Viegas, S.F.; Ferren, E.L.; Self, J.; Tencer, A.F. Comparative mechanical properties of various kirschner wire
configurations in transverse and oblique phalangeal fractures. J. Hand Surg. 1988, 13, 246–253. [CrossRef]
83. Stark, H.H.; Rickard, T.; Zemel, N.; Ashworth, C. Treatment of ununited fractures of the scaphoid by iliac
bone grafts and kirschner-wire fixation. J. Bone Jt. Surg. Am. 1988, 70, 982–991. [CrossRef]
84. Hope, P.G.; Williamson, D.M.; Coates, C.J.; Cole, W.G. Biodegradable pin fixation of elbow fractures in
children. A randomised trial. J. Bone Jt. Surg. Br. Vol. 1991, 73, 965–968. [CrossRef]
85. Leppilahti, J.; Jalovaara, P. Migration of kirschner wires following fixation of the clavicle—A report of 2 cases.
Acta Orthop. Scand. 1999, 70, 517–519. [CrossRef] [PubMed]
86. Gbur, J.L.; Lewandowski, J.J. Fatigue and fracture of wires and cables for biomedical applications. Int. Mater.
Rev. 2016, 61, 231–314. [CrossRef]
87. Chen, Q.; Thouas, G.A. Metallic implant biomaterials. Mater. Sci. Eng. R Rep. 2015, 87, 1–57. [CrossRef]
88. Prasad, K.; Bazaka, O.; Chua, M.; Rochford, M.; Fedrick, L.; Spoor, J.; Symes, R.; Tieppo, M.; Collins, C.;
Cao, A.; et al. Metallic biomaterials: Current challenges and opportunities. Materials 2017, 10, 884. [CrossRef]
[PubMed]
89. Clauss, M.; Graf, S.; Gersbach, S.; Hintermann, B.; Ilchmann, T.; Knupp, M. Material and biofilm load of k
wires in toe surgery: Titanium versus stainless steel. Clin. Orthop. Relat. Res. 2013, 471, 2312–2317. [CrossRef]
[PubMed]
90. Jensen, C.; Jensen, C. Biodegradable pins versus kirschner wires in hand surgery. J. Hand Surg. Br. Eur. Vol.
1996, 21, 507–510. [CrossRef]
91. Plaga, B.; Royster, R.; Donigian, A.; Wright, G.; Caskey, P. Fixation of osteochondral fractures in rabbit
knees. A comparison of kirschner wires, fibrin sealant, and polydioxanone pins. Bone Jt. J. 1992, 74, 292–296.
[CrossRef]
92. Casteleyn, P.P.; Handelberg, F.; Haentjens, P. Biodegradable rods versus kirschner wire fixation of wrist
fractures. A randomised trial. J. Bone Jt. Surg. Br. Vol. 1992, 74, 858–861. [CrossRef]
93. Juutilainen, T.; Patiälä, H.; Rokkanen, P.; Törmälä, P. Biodegradable wire fixation in olecranon and patella
fractures combined with biodegradable screws or plugs and compared with metallic fixation. Arch. Orthop.
Trauma Surg. 1995, 114, 319–323. [CrossRef] [PubMed]
94. Lurate, B.; Mukherjee, D.; Kruse, R.; Albright, J. Fixation of osteochondral fractures with absorbable pins.
In Proceedings of the 1995 Fourteenth Southern Biomedical Engineering Conference, Shreveport, LA, USA,
7–9 April 1995; pp. 57–58.
95. Maruyama, T.; Saha, S.; Mongiano, D.O.; Mudge, K. Metacarpal fracture fixation with absorbable
polyglycolide rods and stainless steel k wires: A biomechanical comparison. J. Biomed. Mater. Res. Part A
1996, 33, 9–12. [CrossRef]
96. Tian, P.; Liu, X. Surface modification of biodegradable magnesium and its alloys for biomedical applications.
Regen. Biomater. 2015, 2, 135–151. [CrossRef] [PubMed]

135
Metals 2018, 8, 212

97. Witte, F.; Hort, N.; Vogt, C.; Cohen, S.; Kainer, K.U.; Willumeit, R.; Feyerabend, F. Degradable biomaterials
based on magnesium corrosion. Curr. Opin. Solid State Mater. Sci. 2008, 12, 63–72. [CrossRef]
98. Hanzi, A.C.; Gerber, I.; Schinhammer, M.; Loffler, J.F.; Uggowitzer, P.J. On the in vitro and in vivo degradation
performance and biological response of new biodegradable Mg-Y-Zn alloys. Acta Biomater. 2010, 6, 1824–1833.
[CrossRef] [PubMed]
99. KubÁSek, J.; VojtĚCh, D. Structural and corrosion characterization of biodegradable Mg–Re (Re = Gd, Y, Nd)
alloys. Trans. Nonferrous Met. Soc. China 2013, 23, 1215–1225. [CrossRef]
100. Chen, L.; Bin, Y.; Zou, W.; Wang, X.; Li, W. The influence of Sr on the microstructure, degradation and stress
corrosion cracking of the Mg alloys—ZK40xSr. J. Mech. Behav. Biomed. Mater. 2017, 66, 187–200. [CrossRef]
[PubMed]
101. Li, Z.; Gu, X.; Lou, S.; Zheng, Y. The development of binary Mg-Ca alloys for use as biodegradable materials
within bone. Biomaterials 2008, 29, 1329–1344. [CrossRef] [PubMed]
102. Gu, X.; Zheng, Y.; Cheng, Y.; Zhong, S.; Xi, T. In vitro corrosion and biocompatibility of binary magnesium
alloys. Biomaterials 2009, 30, 484–498. [CrossRef] [PubMed]
103. Lietaert, K.; Weber, L.; Van Humbeeck, J.; Mortensen, A.; Luyten, J.; Schrooten, J. Open cellular magnesium
alloys for biodegradable orthopaedic implants. J. Magn. Alloys 2013, 1, 303–311. [CrossRef]
104. Mutlu, I.; Oktay, E. Influence of fluoride content of artificial saliva on metal release from 17-4 PH stainless
steel foam for dental implant applications. J. Mater. Sci. Technol. 2013, 29, 582–588. [CrossRef]
105. Cui, W.; Beniash, E.; Gawalt, E.; Xu, Z.; Sfeir, C. Biomimetic coating of magnesium alloy for enhanced
corrosion resistance and calcium phosphate deposition. Acta Biomater. 2013, 9, 8650–8659. [CrossRef]
[PubMed]
106. Erdmann, N.; Angrisani, N.; Reifenrath, J.; Lucas, A.; Thorey, F.; Bormann, D.; Meyer-Lindenberg, A.
Biomechanical testing and degradation analysis of MgCa0.8 alloy screws: A comparative in vivo study in
rabbits. Acta Biomater. 2011, 7, 1421–1428. [CrossRef] [PubMed]
107. Kramer, M.; Schilling, M.; Eifler, R.; Hering, B.; Reifenrath, J.; Besdo, S.; Windhagen, H.; Willbold, E.;
Weizbauer, A. Corrosion behavior, biocompatibility and biomechanical stability of a prototype
magnesium-based biodegradable intramedullary nailing system. Mater. Sci. Eng. C Mater. Biol. Appl.
2016, 59, 129–135. [CrossRef] [PubMed]
108. Chaya, A.; Yoshizawa, S.; Verdelis, K.; Myers, N.; Costello, B.J.; Chou, D.T.; Pal, S.; Maiti, S.; Kumta, P.N.;
Sfeir, C. In vivo study of magnesium plate and screw degradation and bone fracture healing. Acta Biomater.
2015, 18, 262–269. [CrossRef] [PubMed]
109. Mc, B.E. Absorbable metal in bone surgery: A further report on the use of magnesium alloys. J. Am.
Med. Assoc. 1938, 111, 2464–2467.
110. Windhagen, H.; Radtke, K.; Weizbauer, A.; Diekmann, J.; Noll, Y.; Kreimeyer, U.; Schavan, R.;
Stukenborg-Colsman, C.; Waizy, H. Biodegradable magnesium-based screw clinically equivalent to titanium
screw in hallux valgus surgery: Short term results of the first prospective, randomized, controlled clinical
pilot study. Biomed. Eng. Online 2013, 12, 62. [CrossRef] [PubMed]
111. Ezechieli, M.; Ettinger, M.; Konig, C.; Weizbauer, A.; Helmecke, P.; Schavan, R.; Lucas, A.; Windhagen, H.;
Becher, C. Biomechanical characteristics of bioabsorbable magnesium-based (mgyrezr-alloy) interference
screws with different threads. Knee Surg. Sports Traumatol. Arthrosc. 2016, 24, 3976–3981. [CrossRef]
[PubMed]
112. Tie, D.; Feyerabend, F.; Muller, W.D.; Schade, R.; Liefeith, K.; Kainer, K.U.; Willumeit, R. Antibacterial
biodegradable Mg-Ag alloys. Eur. Cells Mater. 2013, 25, 284–298. [CrossRef]
113. Jahn, K.; Saito, H.; Taipaleenmaki, H.; Gasser, A.; Hort, N.; Feyerabend, F.; Schluter, H.; Rueger, J.M.;
Lehmann, W.; Willumeit-Romer, R.; et al. Intramedullary Mg2 Ag nails augment callus formation during
fracture healing in mice. Acta Biomater. 2016, 36, 350–360. [CrossRef] [PubMed]
114. Bian, D.; Zhou, W.; Deng, J.; Liu, Y.; Li, W.; Chu, X.; Xiu, P.; Cai, H.; Kou, Y.; Jiang, B.; et al. Development
of magnesium-based biodegradable metals with dietary trace element germanium as orthopaedic implant
applications. Acta Biomater. 2017, 64, 421–436. [CrossRef] [PubMed]
115. Koo, Y.; Lee, H.B.; Dong, Z.; Kotoka, R.; Sankar, J.; Huang, N.; Yun, Y. The effects of static and dynamic
loading on biodegradable magnesium pins in vitro and in vivo. Sci. Rep. 2017, 7, 14710. [CrossRef] [PubMed]
116. Liu, B.; Zheng, Y.F. Effects of alloying elements (Mn, Co, Al, W, Sn, B, C and S) on biodegradability and
in vitro biocompatibility of pure iron. Acta Biomater. 2011, 7, 1407–1420. [CrossRef] [PubMed]

136
Metals 2018, 8, 212

117. Yang, K.; Zhou, C.; Fan, H.; Fan, Y.; Jiang, Q.; Song, P.; Fan, H.; Chen, Y.; Zhang, X. Bio-functional design,
application and trends in metallic biomaterials. Int. J. Mol. Sci. 2017, 19, 24. [CrossRef] [PubMed]
118. Liu, B.; Zheng, Y.F.; Ruan, L. In vitro investigation of Fe30Mn6Si shape memory alloy as potential
biodegradable metallic material. Mater. Lett. 2011, 65, 540–543. [CrossRef]
119. Hermawan, H.; Dube, D.; Mantovani, D. Developments in metallic biodegradable stents. Acta Biomater. 2010,
6, 1693–1697. [CrossRef] [PubMed]
120. Capek, J.; Msallamova, S.; Jablonska, E.; Lipov, J.; Vojtech, D. A novel high-strength and highly corrosive
biodegradable Fe-Pd alloy: Structural, mechanical and in vitro corrosion and cytotoxicity study. Mater. Sci.
Eng. C Mater. Biol. Appl. 2017, 79, 550–562. [CrossRef] [PubMed]
121. Hufenbach, J.; Wendrock, H.; Kochta, F.; Kühn, U.; Gebert, A. Novel biodegradable Fe-Mn-C-S alloy with
superior mechanical and corrosion properties. Mater. Lett. 2017, 186, 330–333. [CrossRef]
122. Wang, H.; Zheng, Y.; Liu, J.; Jiang, C.; Li, Y. In vitro corrosion properties and cytocompatibility of Fe-Ga
alloys as potential biodegradable metallic materials. Mater. Sci. Eng. C 2017, 71, 60–66. [CrossRef] [PubMed]
123. Babcock, W.W. Metallic sutures and ligatures. Surg. Clin. N. Am. 1947, 27, 1435–1460. [CrossRef]
124. Chellamani, K.; Veerasubramanian, D.; Balaji, R. Surgical Sutures: An Overview. J. Acad. Ind. Res. 2013, 1,
778–782.
125. Seitz, J.-M.; Wulf, E.; Freytag, P.; Bormann, D.; Bach, F.-W. The manufacture of resorbable suture material
from magnesium. Adv. Eng. Mater. 2010, 12, 1099–1105. [CrossRef]
126. Seitz, J.-M.; Utermöhlen, D.; Wulf, E.; Klose, C.; Bach, F.-W. The manufacture of resorbable suture material
from magnesium—Drawing and stranding of thin wires. Adv. Eng. Mater. 2011, 13, 1087–1095. [CrossRef]
127. Bai, J.; Yin, L.; Lu, Y.; Gan, Y.; Xue, F.; Chu, C.; Yan, J.; Yan, K.; Wan, X.; Tang, Z. Preparation, microstructure
and degradation performance of biomedical magnesium alloy fine wires. Prog. Natl. Sci. Mater. Int. 2014, 24,
523–530. [CrossRef]
128. Zhang, S.; Liu, J.; Yan, J.; Chen, Y.; Zhao, C.; Zhang, Y.; Jiang, M.; Xu, H.; Ni, J.; Zhang, X. In vivo degradation
and biocompatibility of linear cutter staples made of high purity magnesium. Eur. Cells Mater. 2014, 28, 76.
129. Wu, H.; Zhao, C.; Ni, J.; Zhang, S.; Liu, J.; Yan, J.; Chen, Y.; Zhang, X. Research of a novel biodegradable
surgical staple made of high purity magnesium. Bioact. Mater. 2016, 1, 122–126. [CrossRef]
130. Cao, J.; Jiang, K.W.; Yang, X.D.; Shen, Z.L.; Guo, P.; Yan, Y.C.; Cui, Y.C.; Han, L.; Lv, Y.; Ye, Y.J.; et al.
[animal experimental study of biodegradable magnesium alloy stapler for gastrointestinal anastomosis].
Zhonghua Wei Chang Wai Ke Za Zhi Chin. J. Gastrointest. Surg. 2013, 16, 772–776.
131. Moore, W.R.; Graves, S.E.; Bain, G.I. Synthetic bone graft substitutes. ANZ J. Surg. 2001, 71, 354–361.
[CrossRef] [PubMed]
132. Zhang, J.; Liu, W.; Schnitzler, V.; Tancret, F.; Bouler, J.-M. Calcium phosphate cements for bone substitution:
Chemistry, handling and mechanical properties. Acta Biomater. 2014, 10, 1035–1049. [CrossRef] [PubMed]
133. Zimmermann, G.; Moghaddam, A. Allograft bone matrix versus synthetic bone graft substitutes. Injury
2011, 42, S16–S21. [CrossRef] [PubMed]
134. Claes, L.; Hoellen, I.; Ignatius, A. Biodegradable bone cements. Der Orthop. 1997, 26, 459–462. [CrossRef]
[PubMed]
135. Geffers, M.; Groll, J.; Gbureck, U. Reinforcement strategies for load-bearing calcium phosphate biocements.
Materials 2015, 8, 2700–2717. [CrossRef]
136. Vaishya, R.; Chauhan, M.; Vaish, A. Bone cement. J. Clin. Orthop. Trauma 2013, 4, 157–163. [CrossRef]
[PubMed]
137. Bohner, M. Design of ceramic-based cements and putties for bone graft substitution. Eur. Cells Mater. 2010,
20, 3–10. [CrossRef]
138. Dos Santos, L.A.; Carrodéguas, R.G.; Boschi, A.O.; Fonseca de Arruda, A.C. Fiber-enriched double-setting
calcium phosphate bone cement. J. Biomed. Mater. Res. Part A 2003, 65, 244–250. [CrossRef] [PubMed]
139. Von Gonten, A.; Kelly, J.; Antonucci, J.M. Load-bearing behavior of a simulated craniofacial structure
fabricated from a hydroxyapatite cement and bioresorbable fiber-mesh. J. Mater. Sci. Mater. Med. 2000, 11,
95–100. [CrossRef] [PubMed]
140. Xu, H.H.K.; Quinn, J.B. Calcium phosphate cement containing resorbable fibers for short-term reinforcement
and macroporosity. Biomaterials 2002, 23, 193–202. [CrossRef]
141. Zhang, Y.; Xu, H.H. Effects of synergistic reinforcement and absorbable fiber strength on hydroxyapatite
bone cement. J. Biomed. Mater. Res. Part A 2005, 75, 832–840. [CrossRef] [PubMed]

137
Metals 2018, 8, 212

142. Pan, Z.; Jiang, P.; Fan, Q.; Ma, B.; Cai, H. Mechanical and biocompatible influences of chitosan fiber and
gelatin on calcium phosphate cement. J. Biomed. Mater. Res. Part B Appl. Biomater. 2007, 82, 246–252.
[CrossRef] [PubMed]
143. Krüger, R.; Groll, J. Fiber reinforced calcium phosphate cements–on the way to degradable load bearing
bone substitutes? Biomaterials 2012, 33, 5887–5900. [CrossRef] [PubMed]
144. Krüger, R.; Seitz, J.-M.; Ewald, A.; Bach, F.-W.; Groll, J. Strong and tough magnesium wire reinforced
phosphate cement composites for load-bearing bone replacement. J. Mech. Behav. Biomed. Mater. 2013, 20,
36–44. [CrossRef] [PubMed]
145. Li, T.; He, Y.; Zhou, J.; Tang, S.; Yang, Y.; Wang, X. Effects of scandium addition on biocompatibility of
biodegradable Mg–1.5Zn–0.6Zr alloy. Mater. Lett. 2018, 215, 200–202. [CrossRef]
146. Salleh, E.M.; Ramakrishnan, S.; Hussain, Z. Synthesis of biodegradable Mg-Zn alloy by mechanical alloying:
Effect of milling time. Procedia Chem. 2016, 19, 525–530. [CrossRef]
147. Gui, Z.; Kang, Z.; Li, Y. Mechanical and corrosion properties of Mg-Gd-Zn-Zr-Mn biodegradable alloy by
hot extrusion. J. Alloys Compd. 2016, 685, 222–230. [CrossRef]
148. Dai, J.; Zhang, X.; Yin, Q.; Ni, S.; Ba, Z.; Wang, Z. Friction and wear behaviors of biodegradable
Mg-6Gd-0.5Zn-0.4Zr alloy under simulated body fluid condition. J. Magn. Alloys 2017, 5, 448–453. [CrossRef]
149. Kubásek, J.; Vojtěch, D.; Jablonská, E.; Pospíšilová, I.; Lipov, J.; Ruml, T. Structure, mechanical characteristics
and in vitro degradation, cytotoxicity, genotoxicity and mutagenicity of novel biodegradable Zn–Mg alloys.
Mater. Sci. Eng. C 2016, 58, 24–35. [CrossRef] [PubMed]
150. Shen, C.; Liu, X.; Fan, B.; Lan, P.; Zhou, F.; Li, X.; Wang, H.; Xiao, X.; Li, L.; Zhao, S. Mechanical
properties, in vitro degradation behavior, hemocompatibility and cytotoxicity evaluation of Zn–1.2Mg
alloy for biodegradable implants. RSC Adv. 2016, 6, 86410–86419. [CrossRef]
151. Dorozhkin, S.V. Calcium orthophosphate coatings on magnesium and its biodegradable alloys. Acta Biomater.
2014, 10, 2919–2934. [CrossRef] [PubMed]
152. Song, Y.W.; Shan, D.Y.; Han, E.H. Electrodeposition of hydroxyapatite coating on AZ91D magnesium alloy
for biomaterial application. Mater. Lett. 2008, 62, 3276–3279. [CrossRef]
153. Rojaee, R.; Fathi, M.; Raeissi, K. Controlling the degradation rate of AZ91 magnesium alloy via sol–gel
derived nanostructured hydroxyapatite coating. Mater. Sci. Eng. C 2013, 33, 3817–3825. [CrossRef] [PubMed]
154. Su, Y.; Li, G.; Lian, J. A chemical conversion hydroxyapatite coating on AZ60 magnesium alloy and its
electrochemical corrosion behaviour. Int. J. Electrochem. Sci. 2012, 7, 11497–11511.
155. Zhang, Y.; Ma, Y.; Chen, M.; Wei, J. Effects of anodizing biodegradable Mg–Zn–Zr alloy on the deposition of
Ca–P coating. Surf. Coat. Technol. 2013, 228, S111–S115. [CrossRef]
156. Tsubakino, H.; Yamamoto, A.; Fukumoto, S.; Watanabe, A.; Sugahara, K.; Inoue, H. High-purity magnesium
coating on magnesium alloys by vapor deposition technique for improving corrosion resistance. Mater. Trans.
2003, 44, 504–510. [CrossRef]
157. Johnson, I.; Akari, K.; Liu, H. Nanostructured hydroxyapatite/poly (lactic-co-glycolic acid) composite coating
for controlling magnesium degradation in simulated body fluid. Nanotechnology 2013, 24, 375103. [CrossRef]
[PubMed]
158. Butev, E.; Esen, Z.; Bor, S. In vitro bioactivity investigation of alkali treated Ti6Al7Nb alloy foams.
Appl. Surf. Sci. 2015, 327, 437–443. [CrossRef]
159. Tang, H.; Wu, T.; Xu, F.; Tao, W.; Jian, X. Fabrication and characterization of Mg(OH)2 films on AZ31
magnesium alloy by alkali treatment. Int. J. Electrochem. Sci. 2017, 12, 1377–1388. [CrossRef]
160. Feng, J.; Chen, Y.; Liu, X.; Liu, T.; Zou, L.; Wang, Y.; Ren, Y.; Fan, Z.; Lv, Y.; Zhang, M. In-situ
hydrothermal crystallization Mg(Oh)2 films on magnesium alloy AZ91 and their corrosion resistance
properties. Mater. Chem. Phys. 2013, 143, 322–329. [CrossRef]
161. Shadanbaz, S.; Dias, G.J. Calcium phosphate coatings on magnesium alloys for biomedical applications:
A review. Acta Biomater. 2012, 8, 20–30. [CrossRef] [PubMed]
162. Harun, W.S.W.; Asri, R.I.M.; Alias, J.; Zulkifli, F.H.; Kadirgama, K.; Ghani, S.A.C.; Shariffuddin, J.H.M.
A comprehensive review of hydroxyapatite-based coatings adhesion on metallic biomaterials. Ceram. Int.
2018, 44, 1250–1268. [CrossRef]
163. Zhao, N.; Zhu, D. Collagen self-assembly on orthopedic magnesium biomaterials surface and subsequent
bone cell attachment. PLoS ONE 2014, 9, e110420. [CrossRef] [PubMed]

138
Metals 2018, 8, 212

164. Wang, Z.-L.; Yan, Y.-H.; Wan, T.; Yang, H. Poly (L-lactic acid)/hydroxyapatite/collagen composite coatings
on AZ31 magnesium alloy for biomedical application. Proc. Inst. Mech. Eng. Part H J. Eng. Med. 2013, 227,
1094–1103. [CrossRef] [PubMed]
165. Chen, C.-Z.; Shi, X.-H.; Zhang, P.-C.; Bai, B.; Leng, Y.-X.; Huang, N. The microstructure and properties of
commercial pure iron modified by plasma nitriding. Solid State Ion. 2008, 179, 971–974. [CrossRef]
166. Zhu, S.; Huang, N.; Xu, L.; Zhang, Y.; Liu, H.; Lei, Y.; Sun, H.; Yao, Y. Biocompatibility of Fe–O films
synthesized by plasma immersion ion implantation and deposition. Surf. Coat. Technol. 2009, 203, 1523–1529.
[CrossRef]
167. Zhou, J.; Yang, Y.; Alonso Frank, M.; Detsch, R.; Boccaccini, A.R.; Virtanen, S. Accelerated degradation
behavior and cytocompatibility of pure iron treated with sandblasting. ACS Appl. Mater. Interfaces 2016, 8,
26482–26492. [CrossRef] [PubMed]
168. Dobatkin, S.V.; Lukyanova, E.A.; Martynenko, N.S.; Anisimova, N.Y.; Kiselevskiy, M.V.; Gorshenkov, M.V.;
Yurchenko, N.Y.; Raab, G.I.; Yusupov, V.S.; Birbilis, N.; et al. Strength, corrosion resistance, and
biocompatibility of ultrafine-grained mg alloys after different modes of severe plastic deformation. IOP Conf.
Ser. Mater. Sci. Eng. 2017, 194, 012004. [CrossRef]

© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

139
metals
Review
The Prospects of Zinc as a Structural Material for
Biodegradable Implants—A Review Paper
Galit Katarivas Levy 1, *, Jeremy Goldman 2 and Eli Aghion 1
1 Department of Materials Engineering, Ben-Gurion University of the Negev, P.O. Box 652,
8410501 Beer-Sheva, Israel; egyon@bgu.ac.il
2 Biomedical Engineering Department, Michigan Technological University, Houghton, MI 49931, USA;
jgoldman@mtu.edu
* Correspondence: levyga@post.bgu.ac.il; Tel.: +972-50-6944074

Received: 31 August 2017; Accepted: 22 September 2017; Published: 1 October 2017

Abstract: In the last decade, iron and magnesium, both pure and alloyed, have been extensively
studied as potential biodegradable metals for medical applications. However, broad experience
with these material systems has uncovered critical limitations in terms of their suitability for clinical
applications. Recently, zinc and zinc-based alloys have been proposed as new additions to the list of
degradable metals and as promising alternatives to magnesium and iron. The main byproduct of
zinc metal corrosion, Zn2+ , is highly regulated within physiological systems and plays a critical role
in numerous fundamental cellular processes. Zn2+ released from an implant may suppress harmful
smooth muscle cells and restenosis in arteries, while stimulating beneficial osteogenesis in bone.
An important limitation of pure zinc as a potential biodegradable structural support, however, lies in
its low strength (σUTS ~30 MPa) and plasticity (ε < 0.25%) that are insufficient for most medical device
applications. Developing high strength and ductility zinc with sufficient hardness, while retaining its
biocompatibility, is one of the main goals of metallurgical engineering. This paper will review and
compare the biocompatibility, corrosion behavior and mechanical properties of pure zinc, as well as
currently researched zinc alloys.

Keywords: zinc; zinc alloys; biodegradable; biocompatible; corrosion degradation; mechanical properties

1. Introduction
Over the last 4 decades, innovations in biomaterials and medical technology have attracted
remarkable attention for their potential to improve human life, by replacing and repairing soft and
hard tissues, such as bone, cartilage, blood vessels, or even entire organs [1,2]. During this time, metals
have become widely used as orthopedic implants, cardiovascular interventional devices, and tissue
engineering scaffolds, due to their high strength and toughness compared with polymers and ceramic
materials [3,4]. Traditional metallic biomaterials with high corrosion resistance, such as titanium alloys,
stainless steels, cobalt–chromium alloys and tantalum are generally used as permanent implants
in patients [3–5]. In most applications, however, the function of the implant is temporary and no
longer needed after full recovery of the treated site. Furthermore, the permanent presence of the
implant can lead to chronic deleterious effects. For instance, metal ions can be released from implanted
devices due to defects in the surface oxide film, eventually resulting in implant fracture. In other cases,
a chronic inflammatory response against the implant may undermine the therapeutic function of the
device. In such circumstances, a second operation may be necessary to extract the implant, resulting in
additional injury and expense [1,3,6,7].
Biodegradable metals represent an alternative approach to the traditional paradigm of corrosion
resistant metals [1,8]. Biodegradable metals are expected to corrode gradually and harmlessly in vivo,
maintain mechanical integrity during the critical tissue healing phase, and then dissolve completely

Metals 2017, 7, 402; doi:10.3390/met7100402 140 www.mdpi.com/journal/metals


Metals 2017, 7, 402

upon fulfilling their mission [8]. In the last decade, iron and magnesium, both pure and alloyed,
have been extensively studied as potential biodegradable metals for medical applications [8–15].
However, broad experience with these material systems has uncovered critical limitations in terms of
their suitability for clinical applications [5,16–22]. For instance, the corrosion rates for Fe and Fe-based
alloys are generally substantially below clinical needs, producing similar problems as found with
permanent implants [8,16,23,24]. Their corrosion products do not appear to be excreted or metabolized
at a satisfactory rate, but rather accumulate and repel neighboring cells and biological matrices,
rather than allowing cells to integrate around and within the original footprint of the degrading
implant [25]. On the other hand, pure and alloyed Mg exhibits insufficient mechanical strength as well
as excessive corrosion rates, accompanied by hydrogen gas evolution, pH increases, and premature
loss of mechanical integrity [7,26–29].
Recently, zinc and zinc-based alloys were proposed as new additions to the list of degradable
metals and as promising alternatives to magnesium and iron [30–34]. The following are advantageous
characteristics of zinc and its alloys for use in medical applications:

• Similar to magnesium and iron, zinc is an essential trace element in the human body. It is
a component of more than 300 enzymes and an even greater number of other proteins, highlighting
its indispensable role in human health. Optimal nucleic acid and protein metabolism, as well
as cell growth, division, and function, require sufficient availability of zinc [30,35]. From this
perspective, zinc ions released from the implant during the degradation phase could integrate
into the normal metabolic activity of the host without producing systemic toxic side effects [36].
• Zinc exhibits high chemical activity, with an electrode potential (−0.762 V) falling between that of
magnesium (−2.372 V) and iron (−0.444 V) [36–38]. Pure zinc metal, therefore, exhibits moderate
degradation rates (faster than the slowly degrading Fe and its alloys, but slower than the rapidly
degrading Mg and its alloys) due to passive layers of moderate stability, formed by corrosion
products [19,39–41].
• Zinc and zinc-based alloys are easier to cast and process due to their low melting points,
low chemical reactivity and good machinability [37,42,43]. For instance, unlike Mg based alloys,
the melting of zinc alloys can more conveniently be performed in air [37].

An important limitation of pure zinc as a potential biodegradable structural support, lies in


its low strength (σUTS ~30 MPa) and plasticity (ε < 0.25%) characteristics that are insufficient for
most medical applications [42,44,45]. Developing zinc with high strength and sufficient hardness is
one of the main goals of metallurgical engineering, to broaden its utility as a biomedical implant.
One of the most powerful tools to improve a metal’s mechanical performance is the addition of
alloying elements to the pure metal matrix [36,44]. Researchers have developed Zn-based alloys
with new and more attractive features, modifying their chemical composition and microstructure,
in order to improve mechanical properties, by principally using solid solution and second-phase
strengthening [45]. Furthermore, improvements in the mechanical properties of pure zinc can be
produced by the thermomechanical refinement of grain size, by extrusion, rolling, etc. [9]. This paper
will review and compare the biocompatibility, corrosion behavior and mechanical properties of
pure zinc as well as currently researched zinc alloys. The various types of zinc alloys, in terms of
compositions and phase constituents, are summarized in Table 1.

141
Metals 2017, 7, 402

Table 1. Common biomedical zinc alloys.

Representive Alloys and Alloying


Family Main Phases References
Elements (wt %)
Zn–0.15Mg α-Zn, Mg2 Zn11 [9]
Zn–0.5Mg α-Zn, Mg2 Zn11 [9]
Zn–1Mg, ZnMg1 α-Zn, Mg2 Zn11 [9,22,42,46,47]
Zn–1.2Mg α-Zn, Mg2 Zn11 [3]
Zn–1.5Mg, ZnMg1.5 α-Zn, Mg2 Zn11 [36,42]
Zn–3Mg, ZnMg3 α-Zn, Mg2 Zn11 [9,42,48]
Zn–1.5Mg–0.1Ca α-Zn, Mg2 Zn11 , CaZn13 [36]
Zn–Mg Zn–1Mg–0.5Ca α-Zn, Mg2 Zn11 , CaZn13 [47]
Zn–1Mg–1Ca α-Zn, Mg2 Zn11 , CaZn13 [44]
Zn–1Mg–0.1Sr Zn, MgZn2 , SrZn13 [49]
Zn–1Mg–0.5Sr Zn, MgZn2 , SrZn13 [49]
Zn–1.5Mg–0.1Sr α-Zn, Mg2 Zn11 , SrZn13 [36]
Zn–1Mg–1Sr α-Zn, Mg2 Zn11 , SrZn13 [44]
Zn–1Mg–0.1Mn Zn, MgZn2 [50]
Zn–1.5Mg–0.1Mn Zn, MgZn2 [50]
Zn–1Ca α-Zn, CaZn13 [22]
Zn–Ca
Zn–1Ca–1Sr α-Zn, CaZn13 , SrZn13 [44]
Zn–Sr Zn–1Sr α-Zn, SrZn13 [22]
Zn–0.5Al Zn, Al [9,51]
Zn–1Al Zn, Al [9,52]
Zn–3Al Zn, Al [52]
Zn–5Al Zn, Al [52]
ZnAl4Cu1 Zn, Al [42]
ZA0.1Mg α-Zn, Mg2 (Zn,Al)11 [51]
ZA0.3Mg α-Zn, Mg2 (Zn,Al)11 [51]
Zn–Al
ZA0.5Mg/Zn-0.5Al-0.5Mg α-Zn, Mg2 (Zn,Al)11 [51,53]
Zn–0.5Al–0.5Mg–0.1Bi Zn, Mg2 (Zn,Al)11 , Mg3 Bi2 [53]
Zn–0.5Al–0.5Mg–0.3Bi Zn, Mg2 (Zn,Al)11 , Mg3 Bi2 [53]
Zn–0.5Al–0.5Mg–0.5Bi Zn, Mg2 (Zn,Al)11 , Mg3 Bi2 [53]
3.5–.5Al, 0.75–1.25Cu, 0.03–0.08Mg Zn, Al [31]
3.5–4.3Al, 2.5–3.2Cu, 0.03–0.06Mg Zn, Al [31]
5.6–6Al, 1.2–1.6Cu Zn, Al [31]
Zn–1Cu η-Zn, ε-CuZn5 [54]
Zn–2Cu η-Zn, ε-CuZn5 [54]
Zn–3Cu η-Zn, ε-CuZn5 [54,55]
Zn–Cu Zn–4Cu η-Zn, ε-CuZn5 [37,54]
Zn–3Cu–0.1Mg Zn, CuZn5 , Mg2 Zn11 [55]
Zn–3Cu–0.5Mg Zn, CuZn5 , Mg2 Zn11 [55]
Zn–3Cu–1Mg Zn, CuZn5 , Mg2 Zn11 [55]
Zn–2Li Zn, α-LiZn4 [25]
Zn–4Li Zn, α-LiZn4 [25]
Zn–Li
Zn–6Li Zn, α-LiZn4 [25]
Zn–Li Zn, α-LiZn4 [56]
Zn–2.5Ag η-Zn, ε-AgZn3 [45]
Zn–Ag Zn–5Ag η-Zn, ε-AgZn3 [45]
Zn–7Ag η-Zn, ε-AgZn3 [45]

2. Zinc in the Human Body


The essential role of zinc in the human body was discovered in 1961 when Iranian farmers
subsisting on a zinc deficiency diet (unrefined flat bread, potatoes, and milk) were found to suffer
from a group of syndromes, consisting of anemia, hypogonadism, and dwarfism. Since this discovery,
interest in the biochemical and clinical aspects of zinc has increased markedly [57]. Presently, it is well
known that zinc is one of the most abundant nutritionally essential elements in the human body [44].
Zinc is present in all organs, tissues, fluids and body secretions, with 86% of its mass residing in
skeletal muscle and bone, 6% in the skin, 5% in the liver, 1.5% in the brain and the remaining
distributed amongst the other tissues [35,58]. At the cellular level, 30–40% is located in the nucleus,

142
Metals 2017, 7, 402

50% in the cytoplasm, organelles and specialized vesicles (for digestive enzymes or hormone storage)
and the remaining portion in the cell membrane [35,44]. The human zinc requirement is estimated
at 15 mg/day [8,10,11,37] and due to its high importance, the body has developed sophisticated
mechanisms to remove zinc from dietary constituents and transport it to desired locations. The human
body is able to absorb zinc from the environment, regulate its concentration in body fluids, transport it
safely to all tissues of the body and to the sites where its presence is required, and safely excrete excess
amounts from the body through the kidneys [35].
As a vital element, zinc plays an important role in numerous physiological systems, including
immune, sexual, neurosensory (cognition and vision), and cell development and growth [19,30].
Zinc participates in numerous fundamental biological functions, such as nucleic acid metabolism,
signal transduction, apoptosis regulation, and gene expression [22]. Zinc is a critical component of
enzymes involved in protein synthesis and energy production. More than 1200 proteins are predicted
to contain, bind, or transport Zn2+ , for example, zinc-finger proteins [44]. At the cellular level, zinc
maintains the structural integrity of biomembranes and is essential for cell proliferation, differentiation
and signaling [19,59]. Zinc plays an important role in bone formation, mineralization, and preservation
of bone mass and can be found in the bone extracellular matrix, where it is co-deposited with calcium
hydroxyapatite [19,22,60]. Indeed, a zinc decrease in bone matrix correlates with aging and skeletal
disease [22,61].
Table 2 summarizes the systemic symptoms resulting from zinc deficiency or excess.
Zinc deficiency contributes to retarded growth, impaired parturition (dystocia), neuropathy, decreased
food intake, diarrhea, dermatitis, hair loss, bleeding tendencies, hypotension, and hypothermia [44].
Zinc deficiency is generally due to insufficient dietary intake. However, it may also be a consequence
of malabsorption and chronic illnesses, such as diabetes, malignancy, liver disease, and sickle cell
disease [35,57]. On the other hand, excessive amounts of Zn2+ in the body may be detrimental to
vital organs, such as the kidney, liver, spleen, brain, and heart [19]. In addition, a prolonged overdose
of zinc results in copper deficiency, provokes hypocupremia, anemia, leucopenia, and neutropenia,
and impairs the Cu–Zn–superoxide dismutase antioxidant enzyme [30,62].

Table 2. The systemic symptoms resulting from zinc deficiency and excess [35,63,64].

Organ or System Zinc Deficiency Zinc Excess


Decreased nerve conduction, neuropsychiatric
Brain Lethargy, focal neuronal deficits.
and neurosensory disorders, mental lethargy
Respiratory disorder after inhalation of zinc
Respiratory tract -
smoke, metal fume fever.
Impaired immune system function, increased
Immune system Altered lymphocyte function.
susceptibility to pathogens
Thymus Thymic atrophy -
Skin lesions, decreased wound healing,
Skin -
acrodermatitis
Gastrointestinal tract - Nausea/vomiting, epigastric pain, diarrhea.
Infertility, retarded genital development,
Reproductive system -
hypogonadism
Prostate - Elevated risk of prostate cancer.

3. Corrosion Behavior of Zinc in the Physiological Environment


The degradation model, proposed by Zheng et al. [8], for biodegradable metals in a neutral
physiological environment, generally occurs via cathodic and anodic reactions. Metallic corrosion
produces hydroxides, oxides, and hydrogen gas by-products [65]. In the specific case of zinc,
as presented in Figure 1, the reactions include a number of intermediate species, according to
the following:
Anodic reaction : 2Zn → 2Zn2+ + 4e− (1)

143
Metals 2017, 7, 402

Cathodic reaction : O2 + 2H2 O + 4e− → 4OH− (2)


2+ −
Zn(OH)2 formation : 2Zn + 4OH → 2Zn(OH)2 (3)

ZnO formation : Zn(OH)2 → ZnO + H2 O (4)

When exposed to body fluid, zinc is oxidized into metal cations following the anodic reaction
in Equation (1). The generated electrons are consumed by a cathodic reaction, corresponding to the
dissolved oxygen reduction in Equation (2). Zn(OH)2 and ZnO corrosion products are likely to form
on the metal surface, without gas evolution, according to Equations (3) and (4). Hence, gas release is
not expected as a result of zinc corrosion, in contrast to the highly problematic hydrogen gas that is
released as a byproduct of magnesium corrosion [8,30,36,59].
It should be noted that the physiological environment is highly aggressive, particularly due to the
high concentration of chloride ions. These ions destabilize the equilibrium between dissolution and
formation of the corrosion product layer, given that chloride ions are able to convert the surface into
soluble chloride salts as follows:

6Zn(OH)2 + Zn2+ + 2Cl− → 6Zn(OH)2 · ZnCl2 (5)

4ZnO + 4H2 O + Zn2+ + 2Cl− → 4Zn(OH)2 · ZnCl2 (6)

The dissolution of the Zn(OH)2 and ZnO surface film components promotes further dissolution
of the exposed metal. Cycles of cathodic and anodic reactions expose the fresh metal substrate
to the physiological solution, form corrosion products, and convert the product into soluble salts.
With progressive exposure, an irregular particle may be separated from the zinc matrix and enter the
surrounding medium [8,36].

Figure 1. Model of the zinc degradation process in physiological fluids.

According to Bowen et al. [59], thin layers of zinc oxide were the only product observed during
early stages (1.5 and 3 months) on the surface of zinc wires (99.99% purity) implanted into rat aorta.
However, as corrosion progressed to 4.5 and 6 months, the corrosion layer thickened and contained three
different phases/layers: calcium phosphate, zinc oxide, and zinc carbonate. The calcium phosphate layer
appeared on the exterior surface without forming a true bulk product. Hence, the calcium phosphate
layer was not thought to play a significant role in zinc biocorrosion. The compact corrosion layer
included the two other phases, zinc oxide, and zinc carbonate, with ZnO appearing in formations
isolated from one another by the zinc carbonate phase. This complex corrosion layer suggests
that corrosion products of zinc in body fluid might be similar to those reported for magnesium.
These ceramic degradation products can accumulate as a function of the local tissue’s physiological
mass transfer rate and therefore may impact tissue healing and remodeling [30,59].

144
Metals 2017, 7, 402

Another important factor that needs to be considered in the corrosion of zinc is the pH value of
the solution. According to Pourbaix diagram, shown in Figure 2, zinc is present as hydrated Zn2+ (aq) ,
over the entire physiological range of pH values and biological standard reduction potentials (~820 mV
to ~−670 mV). The purple arrow in Figure 2 indicates the range of biologically-important standard
potentials determined at pH 7.4 [66]. As reported by Thomas et al. [67] in their investigation of zinc
corrosion as a function of pH, in the pH range 7 to 10, the lowered cathodic reaction rates reduce overall
zinc corrosion rates, and the surface oxides thermodynamically predicted to form in this pH range do
not form an effective corrosion protection barrier. Hence, zinc metals in physiological environments
with a pH of 7.4 will be dissolved over time, as is required for biodegradable medical implants.

Figure 2. Pourbaix diagram of zinc. The purple arrow shows the range of biological standard reduction
potentials at pH 7.4: from ~820 mV to ~−670 mV [66].

4. Biocompatibility and Biological Performances of Zinc and Evaluated Zinc Alloys

4.1. Pure Zinc


After implantation, Zn2+ will be released from the zinc implant to the surrounding extracellular
space and eventually into the bloodstream. Thus, the cellular responses to high extracellular Zn2+
will impact the healing process and biocompatibility with zinc-bearing implants [16]. Furthermore,
the response of the host’s immune system will determine whether the implant becomes biointegrated
and continues to function as designed, or encapsulated in a dense fibrous tissue that may compromise
the intended function of the implant. The inflammatory cells are sensitive to the implant’s corrosion
behavior and cellular accessibility of the surface product, with more porous and actively corroding
interfaces eliciting more benign inflammatory responses, relative to surfaces that are more resistant to
biocorrosion [20,68]. The inflammatory response can therefore potentially be regulated by controlling
implant degradation behavior and rates through alloying and processing.

4.1.1. In Vitro Examination


Törne et al. [32] investigated the initial degradation of zinc exposed to simulated (phosphate
buffered saline (PBS), Ringer’s saline solution) and physiological body fluids (human plasma, whole
blood). They found a decrease in the corrosion rate over time when immersed in body fluids and

145
Metals 2017, 7, 402

an increase when placed in simulated body fluids. The passivation film that formed in physiological
body fluids was uniform and was composed of inorganic corrosion products and organic materials
(biomolecules). In contrast, the simulated body fluids promoted localized corrosion with thick porous
products, primarily composed of zinc phosphates and carbonates.
The cellular response of human bone marrow mesenchymal stem cells (hMSC) and human
vascular cells (HCECs, HASMC, HAEC) to Zn2+ was recently investigated [16,26,69,70]. According to
Zhu et al. [70], Zn biomaterial can support hMSC adhesion and proliferation and zinc ions can lead to
enhanced regulation of genes, cell survival/growth and differentiation, extracellular matrix (ECM)
mineralization, and osteogenesis. The stimulation of osteogenesis by ionic zinc has been supported by
numerous studies [71–79]. In the case of human vascular cells, Zn2+ at low concentrations can enhance
cell viability, proliferation, adhesion, spreading, migration, and F-actin and vinculin expression, while
decreasing cell adhesion strength. In contrast, high concentrations of Zn2+ elicited opposite effects.
Gene expression profiles revealed that the most affected functional genes were related to angiogenesis,
inflammation, cell adhesion, vessel tone, and platelet aggregation [26]. Since the Zn2+ concentration is
intrinsically related to the implant degradation rate, a slow corrosion rate with controlled release of
Zn2+ is desirable to maintain a low concentration profile of Zn2+ in the local tissues, in order to benefit
cellular functions [16]. A high concentration of Zn2+ may overwhelm metal divalent ion-dependent
intracellular signaling pathways, stimulate oxidative stress pathways, inducing apoptosis or necrosis,
and generally contribute to negative side effects in cell populations adjacent to the implant.

4.1.2. In Vivo Examination


Bowen et al. studied the in vivo performance of zinc wires (99.99%) over 6 months when
implanted into the abdominal aorta of adult rats [59,80]. They found uniform corrosion with
linearly increasing corrosion rates over the residence time. The corrosion rate after 1.5 months
was below the 0.02 mm/year degradable stent benchmark and increased to 0.05 mm/year (~0.4 and
~0.97 mg/day, respectively) after 6 months, both of which are far below the daily allowance of zinc
(15 mg/day) [59]. In addition, a histological examination indicated excellent biocompatibility with
the arterial tissue as well as tissue regeneration within the original footprint of the degrading implant.
Intriguingly, observations of low cellular density and a distinct lack of smooth muscle cells adjacent to
the implant interface indicates that the Zn2+ ions released from a zinc implant may suppress restenosis
pathways [80]. Yang et al. [81] investigated the degradation of pure zinc stents over one year in a rabbit
abdominal aorta model. They reported an excellent biocompatibility, without severe inflammation,
platelet aggregation, thrombosis formation or obvious intimal hyperplasia. The degradation rates were
matched to the artery healing process. Moreover, the pure zinc stent retained its mechanical integrity
for 6 months and degraded to 41.75 ± 29.72% of its original volume after 12 months of implantation.

4.2. Zinc Alloys


The biocompatibility of absorbable biomaterial components must be considered, given that all
elements of the metal alloy will eventually pass through the human body [10]. Table 3 summarizes
the pathophysiology and toxicology of zinc and the alloying elements used in current zinc alloys.
Table 4 lists the corrosion performance of different zinc alloys, including their in vitro and in vivo
corrosion rates. In Table 4, the Zn–0.5Al–0.5Mg–0.5Bi alloy exhibits the highest corrosion rate
(0.28 mm/year), which for standard implants is still far below the daily allowance of zinc (15 mg/day).
The tabulated data supports the notion that zinc release during the stent degradation process should
be considered safe to human systems, although toxicity at the local level will need to be examined on
a case-by-case basis.
In the ideal design of degradable biomaterials, elements with potential toxicological effects should
be avoided, and their use should be minimized if they cannot be excluded [10]. Since Mg, Ca, Cu, Mn
and Sr are essential for humans, these elements should be the first choices as alloying elements for
biomedical zinc alloys. Li et al. studied the in vitro and in vivo performance of Zn-1X binary alloys

146
Metals 2017, 7, 402

with the nutrient alloying elements, Mg, Ca and Sr [22]. The in vitro results for human morphologies
on the Zn–1X alloy surfaces, and generated pseudopods and extrace umbilical vein endothelial cells
ECV304 and Human osteosarcoma MG63 demonstrated healthy cell llular matrix secretions, compared
to an unhealthy morphology for cells cultured on pure Zn. When experimental Zn–1X pins were
implanted into mouse femurs, new bone formation in the absence of inflammation was observed
around the implantation site, in particular for the Zn–1Sr alloy. The corrosion rates of the Zn–1X pins
were 0.17, 0.19 and 0.22 mm/year for Zn–1Mg, Zn–1Ca and Zn–1Sr pins, respectively, which is far
below the daily allowance of zinc [22]. Even when adding higher amounts of Mg (3 wt %) to pure
zinc, to a concentration of 0.75 mg/mL (Zn: 0.49 ppm and Mg: 10.75 ppm), the Zn–3Mg alloy extract
exhibited acceptable cytotoxic effects on human osteoblasts. Results of the three main parameters of
cell–material interaction, which includes cell health, cell functionality, and inflammatory responses,
demonstrated acceptable cellular toxicity [19]. The addition of Mn to the Zn–Mg system increased
the susceptibility for galvanic micro-cell corrosion due to the higher electrode potential of Mn [50].
The Zn–xCu system exhibited a slightly increased corrosion rate with increasing Cu concentrations,
compared to pure Zn, but the increase was not significantly different among the Zn–xCu alloys. In vitro
testing demonstrated the cytocompatibilty of Zn–xCu alloys with human endothelial cells as well as
an antibacterial property when the Cu concentration was above 2 wt % [54]. The addition of 0.1–1%
Mg to the Zn–3Cu system increased the corrosion rate, as Mg was added [55].
The second choice of alloying element should be those that have been found to improve the
properties of magnesium alloys, including Li and Al. However, given the potential toxicity of Li and
Al, the addition of these two elements should be limited to low weight percent alloys. Moreover,
the degradation rate of zinc alloys containing these elements should match the tolerance or effective
dose range of these elements in the human body. Zhao et al. [25,56] investigated the effect of Li as a
zinc alloying element. They reported the overall quantities of lithium released from a zinc–lithium
implant (with 0.7 wt % of Li) at two orders of magnitude below the daily bodily consumption
allowances. The in vivo implantation of Zn–1Li demonstrated positive biocompatibility, with no
significant differences in serum zinc observed before, and 1–3 months after, implantation. According
to Bowen et al. [52], biodegradation of a Zn–Al stent, with a weight of ~50 mg and a 5 wt % Al
content—assuming bioabsorption of the entire stent in 2 years and a stable corrosion rate—would
result in a daily intake of ~0.003 mg of Al, substantially below the ~10 mg daily intake for an average
person. In their in vivo studies [20,52], Zn–Al (1, 3 and 5% Al) strips were implanted in the wall of the
abdominal aorta of adult Sprague–Dawley rats. The Zn–Al systems exhibited acceptable compatibility
with surrounding arterial tissue, as the histopathological analysis failed to identify necrotic tissue in
the samples examined, although indications of chronic and acute inflammation were both identified.
Moreover, the pattern of corrosion was modified through Al additions, with intergranular corrosion
observed in all Zn–Al alloys. Intergranular corrosion accelerates the oxidation of zinc to zinc oxide,
whose volume expansion produces implant cracking and fragmentation. The addition of 0.1–0.5 wt
% Mg to the Zn–Al system decreased the corrosion rate with increasing Mg content, after 720 h of
immersion in simulated body fluid (SBF) solution. The cytotoxicity results demonstrated acceptable
biocompatibility of the Zn–0.5Al–0.5Mg alloy, with superior antibacterial activity relative to the other
alloys [51].
Presently, the pathophysiology and toxicology of Ag and Bi remain unclear. Silver has been
used clinically for decades to treat burns and assist with wound healing, due to its antibacterial
properties, and has diverse medical applications [45,82,83]. Bi is generally considered less toxic
than other heavy metal elements, such as antimony, and purified bismuth metal has been used to
prepare a number of pharmaceutical products [53,83]. In two studies, the addition of Ag (2.5–7%)
to pure zinc matrix and the addition of Bi (0.1–0.5%) to ternary Zn–Al–Mg alloy slightly increased
the degradation rate. This is likely due to the increased galvanic coupling between the Zn and
AgZn3 /Mg3 Bi2 secondary phases [45,53]. In vitro cytotoxicity tests indicated a more toxic effect on
MC3T3-E1 cells for Zn–Al–Mg–Bi alloys, compared to the Zn–Al–Mg alloy [53].

147
Table 3. Summary of the pathophysiology and toxicology of zinc and select alloying elements [8,10,11,83] and their effect on zinc alloys.

Element Blood Serum Level Daily Allowance Pathophysiology Toxicology Effect on Zinc Alloys
Essential Elements
Activator of many enzymes; co-regulator of
Mg: ↑mechanical properties &
Metals 2017, 7, 402

Mg 17.7–25.8 mg/L 700 mg protein synthesis and muscle contraction; Excessive Mg leads to nausea
↑corrosion rates [22]
stabilizer of DNA and RNA
More than 99% have structural functions in
the skeleton; the solution Ca has signaling Ca: ↑mechanical properties &
Ca 36.8–39.8 mg/L 800 mg Inhibit the intestinal absorption of other essential minerals
functions, including muscle contraction, ↑corrosion rates [22]
blood clotting, cell function, etc.
Component of several metalloproteins;
Iron toxicity gives rise to lesions in the gastrointestinal tract, Fe: ↑corrosion rates by galvanic
Fe 5000–17,600 mg/L 10–20 mg crucial in vital biochemical activities, i.e.,
shock and liver damage corrosion mechanism [15]
oxygen sensing and transport
Essential Trace Elements
Trace element; appears in all enzyme Neurotoxic and hinders bone development at higher
Zn 0.8–1.14 mg/L 15 mg -
classes; most Zn appears in muscle concentrations
Cu plays a vital role in the immune system;
Excessive Cu (>1 mg/day) can cause neurodegenerative
has beneficial effects on endothelial cell ↑Cu (1–4%): ↑mechanical properties
Cu 4.51–8.32 mg/L 1–3 mg diseases, including Alzheimer’s, Menkes and
proliferation and has been reported to & ↑corrosion rates [37,54]
Wilson’s diseases [55]
enhance antibacterial properties [55]
Activator of enzymes; Mn deficiency is Mn improves the casting process.

148
Mn <0.0008 mg/L 4 mg related to osteoporosis, diabetes mellitus, Excessive Mn results in neurotoxicity Mn ↑ susceptibility of galvanic
and atherosclerosis micro-cell corrosion [50]
Other Elements
99% is located in bone; shows dose
a Sr: ↑mechanical properties &
Sr 0.17 mg 2 mg dependent metabolic effects on bone; High doses induce skeletal abnormalities
↑corrosion rates [22]
low doses stimulate new bone formation
Used in the treatment of manic Plasma concentrations of 2 mM are associated with reduced Li: ↑ultimate tensile strength,
Li 0.002–0.004 mg/L 0.2–0.6 mg
depressive psychoses kidney function and neurotoxicity, 4 mM may be fatal ↓ductility & ↓corrosion rate [56]
Primarily accumulates in the bone and nervous systems;
Al: ↑mechanical properties &
Al 0.0021–0.0048 mg/L - - implicated in the pathogenesis of Alzheimer’s disease;
↑corrosion rates
can cause muscle fiber damage; decreases osteoblast viability
a Sr concentration in total blood [8].
Table 4. In vitro and in vivo corrosion rates of different zinc alloys.

In Vitro Corrosion Rate In Vivo Corrosion Rate


Alloy References
Electrochemical (μA/cm2 ) Immersion (mm/year) (mm/year)
1.8–9.2 (Hank’s) 0.05 (plasma)
Metals 2017, 7, 402

Zn 0.04 (whole blood) 0.027–0.13 (Hank’s) 0.02–0.05 (1.5–6 months) [9,20,22,23,31,32,45,53,59]


0.035 (PBS)
Zn * 8.98 - - [9]
Zn–Mg
Zn–0.15Mg c 11.52 (Hank’s) 0.17 (Hank’s) - [9]
Zn–0.15Mg *,c 10.98 (Hank’s) - - [9]
Zn–0.5Mg c 11.73 (Hank’s) 0.175 (Hank’s) - [9]
Zn–0.5Mg *,c 11.01 (Hank’s) - - [9]
9.9–11.9 (Hank’s) 0.085–0.18 (Hank’s)
Zn–1Mg f /ZnMg1 b 0.28–1.2 (SBF) 0.06–0.28 (SBF) 0.17 [9,22,42,46,47]
0.74 (PBS) 0.027 (PBS)
Zn–1Mg * 11.32 (Hank’s) 0.12 (SBF) - [9,46]
Zn–1.2Mg 7.7 (Hank’s) 0.08 (Hank’s) - [3]
Zn–1.2Mg * 12.4 (Hank’s) 0.11 (Hank’s) - [3]
Zn–1.5Mg d /ZnMg1.5 0.063 (Hank’s)
b 8.8 (SBF) - [4,36]
0.05 (SBF)

149
b 9.01 (Hank’s) 0.13 (Hank’s)
Zn–3Mg/ZnMg3 - [9,42,48]
7.4 (SBF) 0.06–0.21 (SBF)
Zn–3Mg * 8.6 (Hank’s) - - [9]
Zn–3Mg *** - 0.13 (SBF) - [48]
Zn–1.5Mg–0.1Ca d - 0.12 (Hank’s) - [36]
Zn–1Mg–0.5Ca 4.3 (PBS) 0.37 (PBS) [47]
Zn—1Mg–1Ca e 0.17 (Hank’s) 0.09 (Hank’s) - [44]
Zn—1Mg–0.1Sr 7.85 (Hank’s) - - [49]
Zn–1Mg–0.5Sr 7.83 (Hank’s) - - [49]
Zn–1.5Mg–0.1Sr d - 0.1 (Hank’s) - [36]
Zn–1Mg–1Sr e 0.175 (Hank’s) 0.095 (Hank’s) - [44]
Zn–1Mg–0.1Mn 17.21 (Hank’s) 0.12 (Hank’s) - [50]
Zn–1.5Mg–0.1Mn 9.34 (Hank’s) 0.09 (Hank’s) - [50]
Zn–Ca
Zn–1Ca f 10.75 (Hank’s) 0.09 (Hank’s) 0.19 [22]
Zn–1Ca–1Sr e 0.185 (Hank’s) 0.11 (Hank’s) - [44]
Zn–Sr
Zn–1Sr f 11.76 (Hank’s) 0.095 (Hank’s) 0.22 [22]
Zn–Al
11.08 (Hank’s) 0.14 (Hank’s)
Zn–0.5Al - [9,51]
20 (SBF) 15 (SBF)
Zn–0.5Al *,c 9.6 (Hank’s) - - [9]
Zn–1Al 11.11 (Hank’s) 0.16 (Hank’s) - [9]
Table 4. Cont.

In Vitro Corrosion Rate In Vivo Corrosion Rate


Alloy References
Electrochemical (μA/cm2 ) Immersion (mm/year) (mm/year)
Zn–1Al * 9.7 (Hank’s) - - [9]
Metals 2017, 7, 402

ZnAl4Cu1 b 5.2 (SBF) 0.07 (SBF) - [42]


ZA0.1Mg 17 (SBF) 0.13 (SBF) - [51]
ZA0.3Mg 11.2 (SBF) 0.11 (SBF) - [51]
ZA0.5Mg/Zn–0.5Al–0.5Mg 9.5 (SBF) 0.11–0.15 (SBF) - [51,53]
Zn–0.5Al–0.5Mg–0.1Bi 12 (SBF) 0.17 (SBF) - [53]
Zn–0.5Al–0.5Mg–0.3Bi 16 (SBF) 0.2 (SBF) - [53]
Zn–0.5Al–0.5Mg–0.5Bi 23 (SBF) 0.28 (SBF) - [53]
ZA4–1 2.986 (Hank’s) - - [31]
ZA4–3 7.209 (Hank’s) - - [31]
ZA6–1 5.331 (Hank’s) - - [31]
Zn–Cu
Zn–1Cu *,a - 0.033 (c-SBF) - [54]
Zn–2Cu *,a - 0.027 (c-SBF) - [54]
0.012 (Hank’s)
Zn–3Cu *,a 0.372 (Hank’s) - [54,55]
0.03 (c-SBF)
0.009 (Hank’s)
Zn–4Cu *,a 4.1 (Hank’s) - [37,54]

150
0.025 (c-SBF)
Zn–3Cu–0.1Mg * 1.18 (Hank’s) 0.023 (Hank’s) - [55]
Zn–3Cu–0.5Mg * 1.56 (Hank’s) 0.03 (Hank’s) - [55]
Zn–3Cu–1Mg * 12.4 (Hank’s) 0.0432 (Hank’s) - [55]
Zn–Ag
Zn–2.5Ag 9.2 (Hank’s) 0.079 (Hank’s) - [45]
Zn–5Ag 9.7 (Hank’s) 0.081 (Hank’s) - [45]
Zn–7Ag 9.9 (Hank’s) 0.084 (Hank’s) - [45]
Zn–Li
Zn–2Li 0.011 (SBF) - - [25]
Zn–4Li 0.004 (SBF) - - [25]
Zn–6Li 0.0038 (SBF) - - [25]
Zn–1Li - - 0.02–0.05 [56]
SBF: simulated body fluid; PBS: phosphate buffered saline; * Hot extrusion; *** Homogenisation; a Data gathered from figure in literature [54] (Figure 7); b Data gathered from figure in
literature [42] (Figure 5b); c Data gathered from figure in literature [9] (Figure 12); d Data gathered from figure in literature [36] (Figure 5); e Data gathered from figure in literature [44]
(Figure 7); f Data gathered from figure in literature [22] (Figure 3b).
Metals 2017, 7, 402

5. Mechanical Properties of Zinc Alloys


Implanted biomaterials, such as bone plates and stents, should be compatible with the mechanical
properties of the substituted tissue [10]. Table 5 compares the mechanical properties of zinc and current
zinc-based alloys to bone and arterial tissues. It is apparent that the mechanical properties of pure zinc
are insufficient for bone and arterial medical device applications, in particular compared to cortical
bone. The most effective way to address this issue is by adding alloying elements and/or through the
refinement of grain size by thermomechanical processing [9].
Zinc alloys exhibit a wide range of ultimate tensile strengths and elongations, from 87 to 399 MPa
and from 0.9% to ~170%, respectively. It can be seen from Table 5 that even minor alloying can
significantly improve mechanical properties; for example, adding 0.15% Mg to pure zinc improves
its ultimate tensile strength from 18 MPa to 250 MPa, and the elongation fraction from 0.32% to
22% [9,44]. Moreover, hot rolling and hot extrusion contribute to the strength and ductility of zinc
alloys [3,9,44,46]. For example: the yield strength (YS), ultimate tensile strength (UTS) and elongation
of as-cast Zn–1Mg–1Ca are 80 MPa, 130 MPa, and 1%, respectively. Meanwhile, the YS, UTS and
elongation properties of Zn-based ternary alloy samples are improved to 138 MPa, 197 MPa, and 8.5%
after hot rolling and 205 MPa, 250 MPa, and 5.2% after hot extrusion, respectively [44]. Therefore, it is
feasible to satisfy strength requirements for zinc alloys through conventional metallurgical approaches.

Table 5. Mechanical properties of bone and arterial tissues, compared with current biomedical
zinc alloys.

Mechanical Properties
Ultimate
Tissue/Alloy Yield Strength Tensile References
Elongation (%) Hardness (HV)
(YS) (MPa) Strength
(UTS) (MPa)
Cortical bone 104.9–114.3 35–283 5–23 - [10,11]
Cancellous bone - 1.5–38 - - [10,11]
Arterial wall - 0.5–1.72 - - [10,11]
Zn c 10 18 0.32 38 [44]
Zn *,c 35 60 3.5 - [44]
Zn **,c,d 30–110 50–140 5.8–36 39 [44,52]
Zn–Mg
Zn—0.15Mg * 114 250 22 52 [9]
Zn–0.5Mg * 159 297 13 65 [9]
Zn–1Mg/ZnMg1 180 340 6 75–86 [9,22,42,47]
Zn–1Mg *,f 175 250 12 - [46]
Zn–1.2Mg 116 130 1.4 93 [3]
Zn–1.2Mg * 220 362 21 96 [3]
Zn-1.5Mg 112 150 1.3 155 [36]
Zn–3Mg/ZnMg3 - 104 2.3 201 [48]
Zn–3Mg * 291 399 1 117 [9]
Zn–3Mg *** - 88 8.8 175 [48]
Zn–1.5Mg–0.1Ca 173 241 1.72 150 [36]
Zn–1Mg–0.5Ca - 150 1.34 116 [47]
Zn–1Mg–1Ca c 80 130 1 90 [44]
Zn–1Mg–1Ca *,c 205 250 5.2 - [44]
Zn–1Mg–1Ca **,c 138 197 8.5 105 [44]
Zn–1Mg–0.1Sr 109 133 1.4 94 [49]
Zn–1Mg–0.5Sr 129 144 1.1 109 [49]
Zn–1.5Mg–0.1Sr 130 209 2.0 145 [36]
Zn–1Mg–1Sr c 85 135 1.2 85 [44]
Zn–1Mg–1Sr *,c 200 250 7.3 - [44]
Zn–1Mg–1Sr **,c 140 200 9.7 90 [44]
Zn–1Mg–0.1Mn 114 132 98 1.11 [50]
Zn–1.5Mg–0.1Mn 115 122 149 0.77 [50]

151
Metals 2017, 7, 402

Table 5. Cont.

Mechanical Properties
Ultimate
Tissue/Alloy Yield Strength Tensile References
Elongation (%) Hardness (HV)
(YS) (MPa) Strength
(UTS) (MPa)
Zn–Ca
Zn–1Ca 119 165 2 73 [22]
Zn–1Ca–1Sr c 83 140 1.1 90 [44]
Zn–1Ca–1Sr *,c 210 260 6.8 - [44]
Zn–1Ca–1Sr **,c 145 203 8.6 85 [44]
Zn–Sr
Zn–1Sr 120 171 2 61 [22]
Zn–Al
Zn–0.5Al * 119 203 33 59 [9]
Zn–1Al * 134 223 24 73 [9]
Zn–1Al **,d 190 220 24 - [52]
Zn–3Al **,d 200 240 30 - [52]
Zn–5Al **,d 240 300 16 - [52]
ZnAl4Cu1 171 210 1 80 [42]
ZA0.1Mg - 87 1.6 79 [51]
ZA0.3Mg - 93 1.7 89 [51]
ZA0.5Mg/Zn–0.5Al–0.5Mg - 92–102 1.73–2.1 94 [51,53]
Zn–0.5Al–0.5Mg–0.1Bi - 102 2.4 102 [53]
Zn–0.5Al–0.5Mg–0.3Bi - 108 2.7 109 [53]
Zn–0.5Al–0.5Mg–0.5Bi - 98 1.97 99 [53]
ZA4–1 75 180 ~112 50 [31]
ZA4–3 110 200 ~130 55 [31]
ZA6–1 175 275 ~170 65 [31]
Zn–Cu
Zn–1Cu * 149 186 21 - [54]
Zn–2Cu * 199 240 46 - [54]
Zn–3Cu * 213 257 47 - [54,55]
Zn–4Cu * 227–250 270 51 - [37,54]
Zn–3Cu–0.1Mg *,b 340 355 5 - [55]
Zn–3Cu–0.5Mg *,b 390 400 2 - [55]
Zn–3Cu–1Mg *,b 427 441 0.9 - [55]
Zn–Ag
Zn–2.5Ag *,a 174 200 35 - [45]
Zn–5Ag *,a 236 250 36 - [45]
Zn–7Ag *,a 258 287 32 - [45]
Zn–Li
Zn–2Li **,e 240 360 14.2 98 [25]
Zn–4Li **,e 420 440 13.7 115 [25]
Zn–6Li **,e 470 560 2.2 136 [25]
Zn–1Li 238 274 17 97 [56]
* Hot extrusion; ** Hot rolling; *** Homogenisation; a Data gathered from figure in literature [45]; (Figure 8a); b Data
gathered from figure in literature [55] (Figure 5b); c Data gathered from figure in literature [44] (Figures 3 and 4);
d Data gathered from figure in literature [52] (Figure 7); e Data gathered from figure in literature [25] (Figure 7);
f Data gathered from figure in literature [46] (Figure 8).

6. Concluding Remarks and Perspectives


The previous several years have seen rapid growth in the research and development of zinc-based
medical devices, due to their biological and biodegradability properties that match the human body.
Zn is one of the most abundant essential elements in the human body, playing essential roles in human
health. Moreover, zinc overcomes the limitations inherent to iron and magnesium, both pure and
alloyed. This includes more suitable corrosion rates as well as easier casting and processing. In this
review, current research progress for zinc and zinc-based alloys has been presented and discussed.

152
Metals 2017, 7, 402

In the case of pure zinc, in vivo studies have demonstrated a great potential for use as biodegradable
stents. The degradation of pure zinc stents proceeded with excellent biocompatibility to local cells and
tissue and was aligned with the time course of arterial healing. Although the pure zinc stent poorly
retained its mechanical integrity during the healing phase, this can be overcome by using zinc alloys
with superior strength.
Despite the advantages, the use of pure Zn as a biodegradable metal is limited due to its
insufficient strength, plasticity and hardness for most medical applications. Adding alloying elements
and refining grain sizes via thermomechanical processing are commonly applied to modify the
mechanical properties of metallic materials. A number of zinc alloys have been developed with
nontoxic and biocompatible alloying elements that have achieved suitable mechanical properties to
serve as structure support for arteries or bone, with promising preliminary results in cell culture and
small animal models. For instance, Zn–Mg and Zn–Al at alloying concentrations of less than one
percent, have enhanced mechanical properties and achieved adequate strength and ductility, without
using excessive quantities of potentially toxic alloying elements. The mechanical properties of these
materials are generally enhanced considerably from the as-cast state, following extrusion to break
up embrittling intermetallics and to refine the grain size. Because each biomedical device (e.g., stent,
bone plate or screw, etc.) requires unique processing conditions that can dramatically change material
properties, it is crucial to evaluate candidate materials through these processing steps.
Looking ahead, researchers are required to translate the wealth of biomedical research data into
an understanding of how these materials will behave within test subjects. Hence, it will be important
to clarify the molecular mechanisms that stimulate beneficial cellular remodeling activities in response
to Zn2+ and the alloying element ions. Next generation zinc implants will need to have tailored
corrosion rates in order to optimize favorable cellular responses and minimize toxicity and negative
inflammatory reactions, specific to the host tissue.

Author Contributions: All the authors wrote the paper.


Conflicts of Interest: The authors declare no conflict of interest.

References
1. Katarivas Levy, G.; Aghion, E. Influence of heat treatment temperature on corrosion characteristics of
biodegradable EW10X04 Mg alloy coated with Nd. Adv. Eng. Mater. 2016, 18, 269–276. [CrossRef]
2. Holzapfel, B.M.; Reichert, J.C.; Schantz, J.-T.; Gbureck, U.; Rackwitz, L.; Nöth, U.; Jakob, F.; Rudert, M.;
Groll, J.; Hutmacher, D.W. How smart do biomaterials need to be? A translational science and clinical point
of view. Adv. Drug Deliv. Rev. 2013, 65, 581–603. [CrossRef] [PubMed]
3. Shen, C.; Liu, X.; Fan, B.; Lan, P.; Zhou, F.; Li, X.; Wang, H.; Xiao, X.; Li, L.; Zhao, S. Mechanical properties,
in vitro degradation behavior, hemocompatibility and cytotoxicity evaluation of Zn–1.2 Mg alloy for
biodegradable implants. RSC Adv. 2016, 6, 86410–86419. [CrossRef]
4. Niinomi, M. Metallic biomaterials. J. Artif. Organs 2008, 11, 105–110. [CrossRef] [PubMed]
5. Katarivas Levy, G.; Ventura, Y.; Goldman, J.; Vago, R.; Aghion, E. Cytotoxic characteristics of biodegradable
EW10X04 Mg alloy after nd coating and subsequent heat treatment. Mater. Sci. Eng. C 2016, 62, 752–761.
[CrossRef] [PubMed]
6. Pospíšilová, I.; Soukupová, V.; Vojtěch, D. Influence of calcium on the structure and mechanical properties of
biodegradable zinc alloys. In Materials Science Forum; Trans Tech Publication Ltd.: Stafa-Zurich, Switzerland, 2017;
pp. 400–403.
7. Aghion, E.; Levy, G.; Ovadia, S. In Vivo behavior of biodegradable Mg–Nd–Y–Zr–Ca alloy. J. Mater. Sci.
2012, 23, 805–812. [CrossRef] [PubMed]
8. Zheng, Y.; Gu, X.; Witte, F. Biodegradable metals. Mater. Sci. Eng. C 2014, 77, 1–34. [CrossRef]
9. Mostaed, E.; Sikora-Jasinska, M.; Mostaed, A.; Loffredo, S.; Demir, A.; Previtali, B.; Mantovani, D.;
Beanland, R.; Vedani, M. Novel zn-based alloys for biodegradable stent applications: Design, development
and in vitro degradation. J. Mech. Behav. Biomed. Mater. 2016, 60, 581–602. [CrossRef] [PubMed]

153
Metals 2017, 7, 402

10. Gu, X.-N.; Zheng, Y.-F. A review on magnesium alloys as biodegradable materials. Front. Mater. Sci. China
2010, 4, 111–115. [CrossRef]
11. Witte, F.; Hort, N.; Vogt, C.; Cohen, S.; Kainer, K.U.; Willumeit, R.; Feyerabend, F. Degradable biomaterials
based on magnesium corrosion. Curr. Opin. Solid State Mater. Sci. 2008, 12, 63–72. [CrossRef]
12. Jablonská, E.; Vojtěch, D.; Fousová, M.; Kubásek, J.; Lipov, J.; Fojt, J.; Ruml, T. Influence of surface
pre-treatment on the cytocompatibility of a novel biodegradable ZnMg alloy. Mater. Sci. Eng. C 2016,
68, 198–204. [CrossRef] [PubMed]
13. Demir, A.G.; Monguzzi, L.; Previtali, B. Selective laser melting of pure Zn with high density for biodegradable
implant manufacturing. Addit. Manuf. 2017, 15, 20–28. [CrossRef]
14. Wang, C.; Yu, Z.; Cui, Y.; Zhang, Y.; Yu, S.; Qu, G.; Gong, H. Processing of a novel Zn alloy micro-tube for
biodegradable vascular stent application. J. Mater. Sci. Technol. 2016, 32, 925–929. [CrossRef]
15. Luo, M.; Shen, W.; Wang, Y.; Allen, M. In Vitro Degradation of Biodegradable Metal Zn And Zn/Fe-Couples
and Their Application as Conductors in Biodegradable Sensors. In Proceedings of the 2015 Transducers-2015
18th International Conference on Solid-State Sensors, Actuators and Microsystems (TRANSDUCERS),
Anchorage, AK, USA, 21–25 June 2015; pp. 1370–1373.
16. Ma, J.; Zhao, N.; Zhu, D. Endothelial cellular responses to biodegradable metal zinc. ACS Biomater. Sci. Eng.
2015, 1, 1174–1182. [CrossRef] [PubMed]
17. Aghion, E.; Levy, G. The effect of Ca on the in vitro corrosion performance of biodegradable Mg-Nd-Y-Zr
alloy. J. Mater. Sci. 2010, 45, 3096–3101. [CrossRef]
18. Levy, G.; Aghion, E. Effect of diffusion coating of Nd on the corrosion resistance of biodegradable Mg
implants in simulated physiological electrolyte. Acta Biomater. 2013, 9, 8624–8630. [CrossRef] [PubMed]
19. Murni, N.; Dambatta, M.; Yeap, S.; Froemming, G.; Hermawan, H. Cytotoxicity evaluation of biodegradable
Zn–3Mg alloy toward normal human osteoblast cells. Mater. Sci. Eng. C 2015, 49, 560–566. [CrossRef]
[PubMed]
20. Guillory, R.J.; Bowen, P.K.; Hopkins, S.P.; Shearier, E.R.; Earley, E.J.; Gillette, A.A.; Aghion, E.; Bocks, M.;
Drelich, J.W.; Goldman, J. Corrosion characteristics dictate the long-term inflammatory profile of degradable
zinc arterial implants. ACS Biomater. Sci. Eng. 2016, 2, 2355–2364. [CrossRef]
21. Hakimi, O.; Ventura, Y.; Goldman, J.; Vago, R.; Aghion, E. Porous biodegradable EW62 medical implants
resist tumor cell growth. Mater. Sci. Eng. C 2016, 61, 516–525. [CrossRef] [PubMed]
22. Li, H.; Xie, X.; Zheng, Y.; Cong, Y.; Zhou, F.; Qiu, K.; Wang, X.; Chen, S.; Huang, L.; Tian, L. Development of
biodegradable Zn-1X binary alloys with nutrient alloying elements Mg, Ca and Sr. Sci. Rep. 2015, 5, 10719.
[CrossRef] [PubMed]
23. Liu, X.; Sun, J.; Yang, Y.; Pu, Z.; Zheng, Y. In vitro investigation of ultra-pure Zn and its mini-tube as potential
bioabsorbable stent material. Mater. Lett. 2015, 161, 53–56. [CrossRef]
24. Kubásek, J.; Vojtěch, D.; Jablonská, E.; Pospíšilová, I.; Lipov, J.; Ruml, T. Structure, mechanical characteristics
and in vitro degradation, cytotoxicity, genotoxicity and mutagenicity of novel biodegradable Zn–Mg alloys.
Mater. Sci. Eng. C 2016, 58, 24–35. [CrossRef] [PubMed]
25. Zhao, S.; McNamara, C.T.; Bowen, P.K.; Verhun, N.; Braykovich, J.P.; Goldman, J.; Drelich, J.W. Structural
characteristics and in vitro biodegradation of a novel Zn-Li alloy prepared by induction melting and hot
rolling. Metall. Mater. Trans. A 2017, 3, 1204–1215. [CrossRef]
26. Ma, J.; Zhao, N.; Zhu, D. Bioabsorbable zinc ion induced biphasic cellular responses in vascular smooth
muscle cells. Sci. Rep. 2016, 6, 26661. [CrossRef] [PubMed]
27. Kubasek, J.; Vojtěch, D. Zn-based alloys as an alternative biodegradable materials. Proc Met. 2012, 5, 23–25.
28. Kubasek, J.; Pospisilova, I.; Vojtech, D.; Jablonska, E.; Ruml, T. Structural, mechanical and cytotoxicity
characterization of as-cast biodegradable Zn–xMg (x = 0.8%–8.3%) alloys. Mater. Tehnol. 2014, 48, 623–629.
29. Dunne, C.F.; Katarivas Levy, G.; Hakimi, O.; Aghion, E.; Twomey, B.; Stanton, K.T. Corrosion behaviour
of biodegradable magnesium alloys with hydroxyapatite coatings. Surf. Coat. Technol. 2016, 289, 37–44.
[CrossRef]
30. Seitz, J.M.; Durisin, M.; Goldman, J.; Drelich, J.W. Recent advances in biodegradable metals for medical
sutures: A critical review. Adv. Healthc. Mater. 2015, 4, 1915–1936. [CrossRef] [PubMed]
31. Wang, C.; Yang, H.; Li, X.; Zheng, Y. In vitro evaluation of the feasibility of commercial Zn alloys as
biodegradable metals. J. Mater. Sci. Technol. 2016, 32, 909–918. [CrossRef]

154
Metals 2017, 7, 402

32. Törne, K.; Larsson, M.; Norlin, A.; Weissenrieder, J. Degradation of zinc in saline solutions, plasma, and
whole blood. J. Biomed. Mater. Res. Part B 2016, 104, 1141–1151. [CrossRef] [PubMed]
33. Bolz, A.; Popp, T. Implantable, Bioresorbable Vessel Wall Support, in Particular Coronary Stent. U.S. Patent
6,287,332 B1, 11 September 2001.
34. Zhou, G.; Gongqi, Q.; Gong, H. Kind of Absorbable High Strength and Toughness Corrosion-Resistant Zinc
Alloy Implant Material for Human Body. U.S. Patent 20170028107 A1, 2 February 2017.
35. Plum, L.M.; Rink, L.; Haase, H. The essential toxin: Impact of zinc on human health. Int. J. Environ. Res.
Public Health 2010, 7, 1342–1365. [CrossRef] [PubMed]
36. Liu, X.; Sun, J.; Qiu, K.; Yang, Y.; Pu, Z.; Li, L.; Zheng, Y. Effects of alloying elements (Ca and Sr) on
microstructure, mechanical property and in vitro corrosion behavior of biodegradable Zn–1.5Mg alloy.
J. Alloys Compd. 2016, 664, 444–452. [CrossRef]
37. Niu, J.; Tang, Z.; Huang, H.; Pei, J.; Zhang, H.; Yuan, G.; Ding, W. Research on a Zn-Cu alloy as
a biodegradable material for potential vascular stents application. Mater. Sci. Eng. C 2016, 69, 407–413.
[CrossRef] [PubMed]
38. Huang, T.; Zheng, Y.; Han, Y. Accelerating degradation rate of pure iron by zinc ion implantation.
Regen. Biomater. 2016, 3, 205–215. [CrossRef] [PubMed]
39. Vojtech, D.; Pospisilova, I.; Michalcova, A.; Maixner, J. Microstructure and mechanical properties of the
micrograined hypoeutectic zn-mg alloy. Int. J. Miner. Metall. Mater. 2016, 23, 1167–1176.
40. Pospíšilová, I.; Vojtěch, D. Zinc alloys for biodegradable medical implants. In Materials Science Forum;
Trans Tech Publlication: Stafa-Zurich, Switzerland, 2014; pp. 457–460.
41. Zhao, L.; Zhang, Z.; Song, Y.; Liu, S.; Qi, Y.; Wang, X.; Wang, Q.; Cui, C. Mechanical properties and in vitro
biodegradation of newly developed porous Zn scaffolds for biomedical applications. Mater. Des. 2016,
108, 136–144. [CrossRef]
42. Vojtěch, D.; Kubásek, J.; Šerák, J.; Novák, P. Mechanical and corrosion properties of newly developed
biodegradable Zn-based alloys for bone fixation. Acta Biomater. 2011, 7, 3515–3522. [CrossRef] [PubMed]
43. Guleryuz, L.; Ipek, R.; Arıtman, I.; Karaoglu, S. Microstructure and Mechanical Properties of Zn-Mg
Alloys as Implant Materials Manufactured by Powder Metallurgy Method. In AIP Conference Proceedings;
AIP Publishing: Melville, NY, USA, 2017.
44. Li, H.; Yang, H.; Zheng, Y.; Zhou, F.; Qiu, K.; Wang, X. Design and characterizations of novel biodegradable
ternary Zn-based alloys with iia nutrient alloying elements mg, ca and sr. Mater. Des. 2015, 83, 95–102.
[CrossRef]
45. Sikora-Jasinska, M.; Mostaed, E.; Mostaed, A.; Beanland, R.; Mantovani, D.; Vedani, M. Fabrication,
mechanical properties and in vitro degradation behavior of newly developed ZnAg alloys for degradable
implant applications. Mater. Sci. Eng. C 2017, 77, 1170–1181. [CrossRef] [PubMed]
46. Gong, H.; Wang, K.; Strich, R.; Zhou, J.G. In vitro biodegradation behavior, mechanical properties,
and cytotoxicity of biodegradable Zn–Mg alloy. J. Biomed. Mater. Res. Part B 2015, 103, 1632–1640. [CrossRef]
[PubMed]
47. Katarivas Levy, G.L.; Leon, A.; Kafri, A.; Ventura, Y.; Drelich, J.W.; Goldman, J.; Vago, R.; Aghion, E.
Evaluation of biodegradable Zn–1%Mg and Zn–1%Mg–0.5%Ca alloys for biomedical applications. J. Mater.
Sci. Mater. Med. 2017. accepted. [CrossRef]
48. Dambatta, M.; Izman, S.; Kurniawan, D.; Farahany, S.; Yahaya, B.; Hermawan, H. Influence of thermal
treatment on microstructure, mechanical and degradation properties of Zn–3Mg alloy as potential
biodegradable implant material. Mater. Des. 2015, 85, 431–437. [CrossRef]
49. Liu, X.; Sun, J.; Yang, Y.; Zhou, F.; Pu, Z.; Li, L.; Zheng, Y. Microstructure, mechanical properties, in vitro
degradation behavior and hemocompatibility of novel Zn–Mg–Sr alloys as biodegradable metals. Mater. Lett.
2016, 162, 242–245. [CrossRef]
50. Liu, X.; Sun, J.; Zhou, F.; Yang, Y.; Chang, R.; Qiu, K.; Pu, Z.; Li, L.; Zheng, Y. Micro-alloying with Mn in
Zn–Mg alloy for future biodegradable metals application. Mater. Des. 2016, 94, 95–104. [CrossRef]
51. Bakhsheshi-Rad, H.; Hamzah, E.; Low, H.; Kasiri-Asgarani, M.; Farahany, S.; Akbari, E.; Cho, M. Fabrication
of biodegradable Zn-Al-Mg alloy: Mechanical properties, corrosion behavior, cytotoxicity and antibacterial
activities. Mater. Sci. Eng. C 2017, 73, 215–219. [CrossRef] [PubMed]

155
Metals 2017, 7, 402

52. Bowen, P.; Seitz, J.; Guillory, R.; Braykovich, J.; Zhao, F.; Goldman, J.; Drelich, J. Evaluation of wrought Zn-Al
alloys (1, 3, and 5 wt % al) through mechanical and in vivo corrosion testing for stent applications. J. Biomed.
Mater. Res. Part B 2016. [CrossRef]
53. Bakhsheshi-Rad, H.; Hamzah, E.; Low, H.; Cho, M.; Kasiri-Asgarani, M.; Farahany, S.; Mostafa, A.; Medraj, M.
Thermal characteristics, mechanical properties, in vitro degradation and cytotoxicity of novel biodegradable
Zn–Al–Mg and Zn–Al–Mg–xBi alloys. Acta Metall. Sin. (Engl. Lett.) 2017, 30, 201–211. [CrossRef]
54. Tang, Z.; Niu, J.; Huang, H.; Zhang, H.; Pei, J.; Ou, J.; Yuan, G. Potential biodegradable Zn-Cu binary
alloys developed for cardiovascular implant applications. J. Mech. Behav. Biomed. Mater. 2017, 72, 182–191.
[CrossRef] [PubMed]
55. Tang, Z.; Huang, H.; Niu, J.; Zhang, L.; Zhang, H.; Pei, J.; Tan, J.; Yuan, G. Design and characterizations of
novel biodegradable Zn-Cu-Mg alloys for potential biodegradable implants. Mater. Des. 2017, 117, 84–94.
[CrossRef]
56. Zhao, S.; Seitz, J.-M.; Eifler, R.; Maier, H.J.; Guillory, R.J.; Earley, E.J.; Drelich, A.; Goldman, J.; Drelich, J.W.
Zn-Li alloy after extrusion and drawing: Structural, mechanical characterization, and biodegradation in
abdominal aorta of rat. Mater. Sci. Eng. C 2017, 76, 301–312. [CrossRef] [PubMed]
57. Roohani, N.; Hurrell, R.; Kelishadi, R.; Schulin, R. Zinc and its importance for human health: An integrative
review. J. Res. Med. Sci. 2013, 18, 144. [PubMed]
58. Jackson, M. Physiology of zinc: General aspects. In Zinc in Human Biology; Springer: Berlin, Germany, 1989;
pp. 1–14.
59. Bowen, P.K.; Drelich, J.; Goldman, J. Zinc exhibits ideal physiological corrosion behavior for bioabsorbable
stents. Adv. Mater. 2013, 25, 2577–2582. [CrossRef] [PubMed]
60. Moonga, B.S.; Dempster, D.W. Zinc is a potent inhibitor of osteoclastic bone resorption in vitro. J. Bone
Miner. Res. 1995, 10, 453–457. [CrossRef] [PubMed]
61. Yamaguchi, M. Role of zinc in bone formation and bone resorption. J. Trace Elem. Exp. Med. 1998, 11, 119–135.
[CrossRef]
62. Bowen, P.K.; Shearier, E.R.; Zhao, S.; Guillory, R.J.; Zhao, F.; Goldman, J.; Drelich, J.W. Biodegradable metals
for cardiovascular stents: From clinical concerns to recent Zn-alloys. Adv. Healthc. Mater. 2016, 5, 1121–1140.
[CrossRef] [PubMed]
63. Kaltenberg, J.; Plum, L.M.; Ober-Blobaum, J.L.; Honscheid, A.; Rink, L.; Haase, H. Zinc signals promote
IL-2-dependent proliferation of T cells. Eur. J. Immunol. 2010, 40, 1496–1503. [CrossRef] [PubMed]
64. Shankar, A.H.; Prasad, A.S. Zinc and immune function: The biological basis of altered resistance to infection.
Am. J. Clin. Nutr. 1998, 68, 447s–463s. [PubMed]
65. Kaur, G. Biodegradable metals as bioactive materials. In Bioactive Glasses: Potential Biomaterials for Future
Therapy; Springer International Publishing: Cham, Switzerland, 2017.
66. Krezel, A.; Maret, W. The biological inorganic chemistry of zinc ions. Arch. Biochem. Biophys. 2016, 611, 3–19.
[CrossRef] [PubMed]
67. Thomas, S.; Birbilis, N.; Venkatraman, M.S.; Cole, I.S. Corrosion of zinc as a function of pH. J. Sci. Eng.
2012, 68. [CrossRef]
68. Drelich, A.; Zhao, S.; Guillory, R.J.; Drelich, J.W.; Goldman, J. Long-term surveillance of zinc implant in
murine artery: Surprisingly steady biocorrosion rate. Acta Biomater. 2017, 58, 539–549. [CrossRef] [PubMed]
69. Shearier, E.R.; Bowen, P.K.; He, W.; Drelich, A.; Drelich, J.; Goldman, J.; Zhao, F. In vitro cytotoxicity,
adhesion, and proliferation of human vascular cells exposed to zinc. ACS Biomater. Sci. Eng. 2016, 2, 634–642.
[CrossRef] [PubMed]
70. Zhu, D.H.; Su, Y.C.; Young, M.L.; Ma, J.; Zheng, Y.F.; Tang, L.P. Biological responses and mechanisms of
human bone marrow mesenchymal stem cells to Zn and Mg biomaterials. ACS Appl. Mater. Interfaces 2017,
9, 27453–27461. [CrossRef] [PubMed]
71. Qiao, Y.Q.; Zhang, W.J.; Tian, P.; Meng, F.H.; Zhu, H.Q.; Jiang, X.Q.; Liu, X.Y.; Chu, P.K. Stimulation of
bone growth following zinc incorporation into biomaterials. Biomaterials 2014, 35, 6882–6897. [CrossRef]
[PubMed]
72. Al Qaysi, M.; Petrie, A.; Shah, R.; Knowles, J.C. Degradation of zinc containing phosphate-based glass as
a material for orthopedic tissue engineering. J. Mater. Sci. 2016. [CrossRef]

156
Metals 2017, 7, 402

73. Zhong, Z.; Ma, J. Fabrication, characterization, and in vitro study of zinc substituted hydroxyapatite/silk
fibroin composite coatings on titanium for biomedical applications. J. Biomater. Appl. 2017, 32, 399–409.
[CrossRef] [PubMed]
74. Yu, J.M.; Xu, L.Z.; Li, K.; Xie, N.; Xi, Y.H.; Wang, Y.; Zheng, X.B.; Chen, X.S.; Wang, M.Y.; Ye, X.J.
Zinc-modified calcium silicate coatings promote osteogenic differentiation through TGF-β/Smad pathway
and osseointegration in osteopenic rabbits. Sci. Rep. 2017, 7, 3440. [CrossRef] [PubMed]
75. Cama, G.; Nkhwa, S.; Gharibi, B.; Lagazzo, A.; Cabella, R.; Carbone, C.; Dubruel, P.; Haugen, H.; Di Silvio, L.;
Deb, S. The role of new zinc incorporated monetite cements on osteogenic differentiation of human
mesenchymal stem cells. Mater. Sci. Eng. C 2017, 78, 485–494. [CrossRef] [PubMed]
76. Ferreira, E.C.S.; Bortolin, R.H.; Freire-Neto, F.P.; Souza, K.S.C.; Bezerra, J.F.; Ururahy, M.A.G.; Ramos, A.M.O.;
Himelfarb, S.T.; Abreu, B.J.; Didone, T.V.N.; et al. Zinc supplementation reduces RANKL/OPG ratio and
prevents bone architecture alterations in ovariectomized and type 1 diabetic rats. Nutr. Res. 2017, 40, 48–56.
[CrossRef] [PubMed]
77. Yu, W.L.; Sun, T.W.; Qi, C.; Ding, Z.Y.; Zhao, H.K.; Zhao, S.C.; Shi, Z.M.; Zhu, Y.J.; Chen, D.Y.; He, Y.H.
Evaluation of zinc-doped mesoporous hydroxyapatite microspheres for the construction of a novel
biomimetic scaffold optimized for bone augmentation. Int. J. Nanomed. 2017, 12, 2293–2306. [CrossRef]
[PubMed]
78. Yusa, K.; Yamamoto, O.; Takano, H.; Fukuda, M.; Iino, M. Zinc-modified titanium surface enhances osteoblast
differentiation of dental pulp stem cells in vitro. Sci. Rep. 2016, 6. [CrossRef] [PubMed]
79. An, S.F.; Gong, Q.M.; Huang, Y.H. Promotive effect of zinc ions on the vitality, migration, and osteogenic
differentiation of human dental pulp cells. Biol. Trace Elem. Res. 2017, 175, 112–121. [CrossRef] [PubMed]
80. Bowen, P.K.; Guillory, R.J.; Shearier, E.R.; Seitz, J.-M.; Drelich, J.; Bocks, M.; Zhao, F.; Goldman, J. Metallic zinc
exhibits optimal biocompatibility for bioabsorbable endovascular stents. Mater. Sci. Eng. C 2015, 56, 467–472.
[CrossRef] [PubMed]
81. Yang, H.; Wang, C.; Liu, C.; Chen, H.; Wu, Y.; Han, J.; Jia, Z.; Lin, W.; Zhang, D.; Li, W.; et al. Evolution
of the degradation mechanism of pure zinc stent in the one-year study of rabbit abdominal aorta model.
Biomaterials 2017, 145, 92–105. [CrossRef] [PubMed]
82. Lansdown, A.B. Silver in health care: Antimicrobial effects and safety in use. In Biofunctional Textiles and the
Skin; Karger: Basel, Switzerland, 2006.
83. Mertz, W. Trace Elements in Human and Animal Nutrition; Academic Press, INC.: Orlando, FL, USA, 1986.

© 2017 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

157
metals
Article
The Suitability of Zn–1.3%Fe Alloy as a
Biodegradable Implant Material
Alon Kafri 1 , Shira Ovadia 2 , Jeremy Goldman 3, *, Jaroslaw Drelich 4 and Eli Aghion 1
1 Department of Materials Engineering, Ben-Gurion University of the Negev, Beer-Sheva 8410501, Israel;
kafri09@gmail.com (A.K.); egyon@bgu.ac.il (E.A.)
2 Faculty of Health Science, Ben-Gurion University of the Negev, P.O. Box 653, Beer-Sheva 84105, Israel;
shiraov@bgu.ac.il
3 Department of Biomedical Engineering, Michigan Technological University, Houghton, MI 49931, USA
4 Department of Materials Science and Engineering, Michigan Technological University,
Houghton, MI 49931, USA; jwdrelic@mtu.edu
* Correspondence: jgoldman@mtu.edu

Received: 5 February 2018; Accepted: 26 February 2018; Published: 28 February 2018

Abstract: Efforts to develop metallic zinc for biodegradable implants have significantly advanced
following an earlier focus on magnesium (Mg) and iron (Fe). Mg and Fe base alloys experience
an accelerated corrosion rate and harmful corrosion products, respectively. The corrosion rate
of pure Zn, however, may need to be modified from its reported ~20 μm/year penetration rate,
depending upon the intended application. The present study aimed at evaluating the possibility of
using Fe as a relatively cathodic biocompatible alloying element in zinc that can tune the implant
degradation rate via microgalvanic effects. The selected Zn–1.3wt %Fe alloy composition produced
by gravity casting was examined in vitro and in vivo. The in vitro examination included immersion
tests, potentiodynamic polarization and impedance spectroscopy, all in a simulated physiological
environment (phosphate-buffered saline, PBS) at 37 ◦ C. For the in vivo study, two cylindrical disks
(seven millimeters diameter and two millimeters height) were implanted into the back midline of male
Wister rats. The rats were examined post implantation in terms of weight gain and hematological
characteristics, including red blood cell (RBC), hemoglobin (HGB) and white blood cell (WBC) levels.
Following retrieval, specimens were examined for corrosion rate measurements and histological
analysis of subcutaneous tissue in the implant vicinity. In vivo analysis demonstrated that the
Zn–1.3%Fe implant avoided harmful systemic effects. The in vivo and in vitro results indicate that
the Zn–1.3%Fe alloy corrosion rate is significantly increased compared to pure zinc. The relatively
increased degradation of Zn–1.3%Fe was mainly related to microgalvanic effects produced by a
secondary Zn11 Fe phase.

Keywords: zinc; bioabsorbable; biodegradable; implants; encapsulation; in vivo

1. Introduction
The development of bioabsorbable metal implants has drawn major attention over the last
two decades. Early studies mainly focused on pure Fe and Mg and their alloys. Bioabsorbable
implants based on Fe exhibit good biocompatibility and excellent mechanical properties [1–5].
However, they suffer from a harmful mode of corrosion that produces a voluminous iron oxide
layer [6]. The steady accumulation of the iron oxide reduces the lumen cross section, compromises
the integrity of the arterial wall by repelling neighboring tissue and cells [7], and stimulates
inflammation [8]. Bioabsorbable implants based on Mg are attractive mainly due to their excellent
biocompatibility [8–20]. Unfortunately, Mg and its alloys exhibit an excessive corrosion rate associated

Metals 2018, 8, 153; doi:10.3390/met8030153 158 www.mdpi.com/journal/metals


Metals 2018, 8, 153

with a potentially harmful release of hydrogen gas [15,18]. The challenges faced with both Fe and Mg
systems have motivated a search for new metallic materials, among them Zn and Zn-based alloys.
Early in vivo studies with pure zinc arterial implants have reported a promising
biocompatibility [21,22]. The biodegradation rate is near the projected ideal value for vascular
stents, and between that of Mg and Fe. In addition, Zn has important physiological roles, exerting
strong anti-atherogenic properties [23], participating in nucleic acid metabolism, signal transduction,
apoptosis regulation, and gene expression [24,25]. Zinc is not toxic to the human body at the low
levels expected to be released from a typical stent [26]. Our extensive preliminary observations have
confirmed the lack of overt toxicity at the interface between a zinc implant and biological tissue,
despite even 20 months residence time in the body [21,27–29]. However, at high levels zinc can cause
toxicity, manifested in poor growth and anemia [30,31]. To avoid negative responses, the corrosion
rate for zinc-based implants may need to be optimized for biodegradable application. Our previous
studies demonstrated that the corrosion rate of a Zn implant can be reduced in the first several weeks
after deployment into an arterial environment by manipulating the oxide film characteristics [32] or
adding a polymeric layer [33]. In addition, earlier research carried out by the authors [27] suggested
that the low degradation rate of arterial implants made from pure zinc can provoke a long-term
process of inflammation and fibrous encapsulation. Encapsulation can be considered as a major
factor contributing to implant failure, as it isolates the implant from the surrounding tissue and can
potentially cause a discontinuation of the biodegradation process [34,35]. This was supported by
Yue et al. [36] which indicated that too low degradation rate is a challenge for future application of Zn
base alloys as adequate biodegradable implant.
The need to improve the strength of Zn and Zn-based materials, manipulate their corrosion
rate and improve their biocompatibility motivated the search for new Zn-based alloys [37–51].
In order to obtain a higher corrosion rate without compromising biocompatibility, the alloying element
must be biocompatible and have the potential to increase the corrosion rate of the Zn-based alloy.
Since Fe is known as a biocompatible element and the intermetallic phases formed between Zn and
Fe all have higher electrochemical potential [52,53], it was considered a suitable alloying element.
The present study aimed at exploring the prospects of Zn–1.3wt %Fe (denoted as Zn–1.3%Fe) alloy as
a biodegradable implant material in terms of in vitro and in vivo behavior.

2. Experimental Procedure
Ingots of pure Zn and Zn–1.3%Fe alloy were produced by gravity casting in a rectangular steel die
having the following dimensions: 6 × 25 × 4.5 cm. Prior to the casting process, pure zinc bars (99.99%)
and pure iron (99%) powder (−325 mesh) in the desired amount were placed in a graphite crucible
and heated to 750 ◦ C for 3 h in a furnace, with stirring every 30 min. All the test samples produced
from the cast ingot were machined from the central part of the ingot, producing surface roughness
type N5. The chemical compositions of the obtained ingots were determined using an Inductively
Coupled Plasma Optical Emission Spectrometer (ICP-SPECTRO, ARCOS FHS-12, Kelve, Germany)
method, with results shown in Table 1.

Table 1. Chemical composition of pure Zn and Zn–1.3%Fe alloy.

Material System Fe (wt %) Pb (wt %) Al (wt %) Cu (wt %) Cd (wt %)


Pure Zn 0.0054 0.0011 0.0026 0.0007 0.0017
Zn–1.3%Fe 1.31 0.0010 0.0079 0.0006 0.0017

Microstructure examination was carried out using scanning electron microscopy (SEM) with a
JEOL JSM-5600 (JEOL, Tokyo, Japan) equipped with an Energy-dispersive X-ray spectroscopy (EDS)
detector (Thermo Fisher Scientific, Waltham, MA, USA) for spot elemental analysis. Identification
of internal phases was conducted using an X-ray diffractometer RIGAKU-2100H (RIGAKU, Tokyo,

159
Metals 2018, 8, 153

Japan) with Cu-Kα. The diffraction parameter was 40 kV/30 mA and the scanning rate was 2◦ /min.
Indication of mechanical properties was obtained by Vickers hardness measurements performed in a
Zwick/Roell Indentec of Quantarad Technologies, with an applied load of 3 kg. The tensile tests were
performed at room temperature using a universal material test machine (Hounsfield H25KT, Horsham,
PA, USA) at a rate of 0.5mm/min.
The in vitro corrosion behavior was examined by an immersion test for up to 20 days, according
to the ASTM G31-12a standard as well as by electrochemical characterization and stress corrosion
analysis. All the in vitro corrosion tests were carried out in phosphate-buffered saline (PBS) solution at
37 ◦ C with pH levels of ~7.4. The selected corrosion medium and environmental conditions aimed at
simulating the natural physiological environment [54–56]. Monitoring of pH levels was conducted by
daily replacement of the PBS solution. The electrochemical characterization included potentiodynamic
polarization analysis and impedance spectroscopy (EIS) using a Bio-Logic SP-200 potentiostat equipped
with EC-Lab software V11.02. The three-electrode cell used for the electrochemical analysis included
a saturated calomel reference electrode (SCE), a platinum counter electrode, and the tested sample
as the working electrode, with an exposure area of 1cm2 . The scanning rate of the potentiodynamic
polarization analysis was 1 mV/s; the EIS measurements were carried out between 10 kHz and 100 mHz
at 10 mV amplitude over the open circuit potential. Prior to electrochemical testing, the samples were
cleaned in an ultrasonic bath for 5 min, washed with alcohol, and dried in hot air.
The metals used for the in vivo assessment included pure zinc as the biodegradable reference
material, the Zn–1.3%Fe allow, and titanium alloy Ti–6Al–4V as a biostable reference material.
The implant geometry of cylindrical disks of 7 mm diameter and 2 mm height was obtained by
regular machining from cast Zn base ingots and wrought billet in the case of the titanium ally. Prior to
implantation, the disks were ultrasonically cleaned in ethanol for 5 min and then in acetone for 3 min
before final air drying.
All the animal experiments were approved by the Ben-Gurion University of the Negev (BGU)
Committee for the Ethical Care and Use of Laboratory Animals (BGU-IACUC). The experiments were
performed according to the Israel Animal Welfare law (1994) and the NRC Guide for the Care and Use
of Laboratory Animals (2011). BGU’s animal care and use program is approved by the Association for
the Assessment and Accreditation of Laboratory Animal Care International (AAALAC). The in vivo
experiments were carried out in the BGU rodent facility. Nine 250 g male Wistar rats (Envigo, Jerusalem,
Israel) were selected for the in vivo assessment. The rats were divided into 3 groups. The first group
(n = 3) was implanted with pure Zn disks, the second group (n = 3) with the Zn–1.3%Fe alloy, and the
third group (n = 3) with Ti–6Al–4V.
For implantation, the rats were anesthetized with an inhalation anesthesia machine using 3%
isoflurane (Terrel TM Piramal Critical Care, Inc., Bethlehem, PA, USA) in 500 mL/min 100% oxygen.
Two cylindrical disks were implanted aseptically subcutaneously in the midline of the back of the rats,
one between scapulas and another in mid-lumbar area. Post alloy implantation, the rats were placed
on a heating pad until they recovered form anesthesia and were able to ambulate. For postoperative
analgesia, rats received 100 mg/kg Dipyrone (Vitamed Pharmaceutical Industries LTD, Binyamina,
Israel) in the drinking water for 3 days. Rats were monitored daily for surgical wound appearance,
locomotion in the cage, grooming activity and general wellbeing for the first week. This was followed
by a weekly evaluation of body weight, attitude and incision appearance. Blood was collected from
the retro-orbital sinus under isoflurane anesthesia as described above. Whole blood (1 mL) was
collected in EDTA for a complete blood count for red blood cells (RBC), hemoglobin (HGB) and
white blood cells (WBC). Serum (1 mL) was collected for determination of Zn levels. Blood work was
analyzed on the day of implantation and at 4 and 8 weeks post-implantation. The blood biochemical
parameters in terms of RBC, HGB and WBC can indicate abnormal situations, such as a significant
increase in Zn content or infection. At 14 weeks the rats were euthanized with intraperitoneal
150 mg/kg Pentobarbital (CTS Chemical Industries Ltd., Hod Hasharon, Israel) for alloy and tissue
harvest. Tissue from each alloy location was placed in 10% formaldehyde for histology. Removal

160
Metals 2018, 8, 153

of corrosion products to calculate the corrosion rate was preformed according to ASTM G1 using
10%NH4 Cl solution, which allowed for a calculation of the corrosion rate.
The statistical analysis was implemented using a one-way ANOVA to determine the significance
of paired comparisons. A p value < 0.05 was selected for statistical difference between means.

3. Results
The typical microstructure of pure zinc and Zn–1.3%Fe alloy are shown in cross-sectional view
(Figure 1), along with a spot chemical analysis at various points, shown in Table 2. A regular structure
without any second phase is present in pure zinc. The Zn–1.3%Fe alloy contained Zn matrix and an
Fe-rich phase that was dispersed homogenously across the bulk of the alloy. The Fe content within the
Fe-rich phase, as obtained by EDS analysis, was between 7 and 10 wt % corresponding with Fe content
in the Zn-delta phase composition [52].

Figure 1. Microstructure at cross-section of (a) Pure Zn. (b,c) Zn–1.3%Fe alloy, macro and close-up
view respectively.

Table 2. Chemical composition of Zn–1.3%Fe at different points shown in Figure 1b.

El Element (wt %) Tested Area Zn Fe


Point 1 90 ± 1 9.8 ± 0.3
Point 2 100 ± 1 0.13 ± 0.08

The X-ray diffraction analysis (Figure 2) revealed the presence of the two major phases: a pure
Zn and Fe rich phase. Additional precipitating phases in the Zn–1.3%Fe alloy appear to be Delta Zn
(ICDD 045-1184), with a possible contribution of Zeta Zn. This result is in accordance with the EDS
analysis of the Fe-rich phase.

161
Metals 2018, 8, 153

Figure 2. X-ray diffraction analysis of pure Zn and Zn–1.3%Fe alloy.

Hardness tests were conducted to begin evaluating the mechanical properties. The hardness of
pure Zn and Zn–1.3%Fe alloy were 40 ± 3 and 56 ± 2 HV (hardness Vickers), respectively. The increased
Zn–1.3%Fe alloy hardness is attributed to the additional Delta Zn phase.
The mechanical properties of pure Zn and Zn–1.3%Fe alloy in as-cast conditions are shown in
Figure 3. While the yield strength (YS) and ultimate tensile strength (UTS) of Zn–1.3%Fe alloy was
improved compared to pure Zn, the ductility was relatively reduced.

Figure 3. Mechanical properties of pure Zn and Zn–1.3%Fe.

Close-up views of the external surfaces of pure Zn and Zn–1.3%Fe alloy after immersion tests of
10 and 20 days are shown in Figure 4. While a general corrosion attack was observed in both samples,
it was significantly more intense in the Zn–1.3%Fe alloy for both exposure times. The corrosion attack
against pure Zn was relatively mild and did not cover the entire sample surface, the attack against
Zn–1.3%Fe alloy was more severe and covered the entire sample surface.

162
Metals 2018, 8, 153

The calculated corrosion rates for pure Zn and Zn–1.3%Fe alloy, as obtained by the immersion
tests, are shown in Figure 5. The corrosion rate of the Zn–1.3%Fe alloy was nearly twice that of pure
Zn after 10 and 20 days of exposure.

Figure 4. Optical microscopy showing corrosion attack at the surface of pure Zn and Zn–1.3%Fe alloy
after immersion tests in phosphate-buffered saline (PBS) solution at 37 ◦ C. (a,b) pure Zn and Zn–1.3%Fe
alloy after 10 days of exposure. (c,d) pure Zn and Zn–1.3%Fe alloy after 20 days of exposure.

Figure 5. Corrosion rate measurements of pure Zn and Zn–1.3%Fe alloy after immersion tests in PBS
solution at 37 ◦ C.

In order to further investigate the corrosion attack mechanism in Zn–1.3%Fe alloy, close-up cross
sectional views of the corroded area were inspected, as shown in Figure 6. The images revealed a
uniform corrosion attack at the external surface (Figure 6a). This phenomenon is in line with the basic
degradation requirements of biodegradable implants in terms of preserving their mechanical integrity
during the critical period after implantation and undergoing degradation without risk of premature
implant fracture during the mechanical scaffolding phase. At higher magnification (Figure 6b),
intact Zn11 Fe particles are seen in the vicinity of corroded matrix. In accordance with our hypothesis,
the cathodic Zn11 Fe particles (−0.87 Vsce vs. −1.03 Vsce for pure Zn [53]) acted as a cathodic phase

163
Metals 2018, 8, 153

to the Zn matrix, causing accelerated degradation by micro-galvanic effect. This mechanism can
explain the higher corrosion rate of the Zn–1.3%Fe alloy compared to pure Zn as quantified by the
immersion test.

Figure 6. Surface cross section showing corrosion attack of Zn–1.3%Fe alloy after 10 days of exposure:
(a) Macro view. (b) Close-up view.

The electrochemical analysis by potentiodynamic polarization is shown in Figure 7, and the


corresponding Tafel extrapolation measurements are listed in Table 3. The polarization curve of
the Zn–1.3%Fe alloy shifted to higher current densities compared to pure Zn, indicating a reduced
corrosion resistance. Tafel extrapolation obtained from the polarization curves support this observation,
identifying a corrosion rate of 0.013 millimeter per year (mm/y) for Zn–1.3%Fe compared to
0.010 mm/y for the pure Zn sample.

Figure 7. Potentiodynamic polarization curves obtained in PBS solution.

Table 3. Corrosion potential (Ecorr ), corrosion current (Icorr ) and corrosion rate (C.R.) obtained by Tafel
extrapolation from polarization curves shown in Figure 7.

C.R (mm/y) Icorr (μA/cm2 ) Ecorr (V) Parameter Specimen C.R (mm/y)
0.010 0.67 −1.02 Pure Zn 0.010
0.013 0.89 −1.04 Zn–1.3%Fe 0.013

164
Metals 2018, 8, 153

The impedance spectroscopy analysis in terms of impedance modifications after immersion times
of 1 and 48 h in PBS solution are shown by Nyquist plots in Figure 8 and Bode plots in Figure 9.
The relatively larger radius of curvature in the Nyquist diagram of the pure Zn sample indicates that
its corrosion resistance was significantly increased compared to the Zn–1.3%Fe sample. Bode diagrams
reveal that after one hour of exposure to the PBS solution, the corrosion resistance of both samples
are nearly the same (around 10,000 Ohm), while after 48 h of exposure the corrosion resistance of
the Zn–1.3%Fe sample decreased from about 8130 Ohm to 4365 Ohm. The results obtained by the
potentiodynamic polarization and impedance spectroscopy analysis are in accordance with the results
from the immersion tests.

Figure 8. Nyquist plots obtained after immersion in PBS solution for up to 48 h.

Figure 9. Bode plots obtained after immersion in PBS solution for up to 48 h.

The in vivo evaluation demonstrated a normal weight gain for all the rats irrespective of implant
type Zn, Figure 10. Post-implantation behavior, cage locomotion and surgical wound appearance were
all normal, with no signs of compromised health.
Zinc serum levels and blood biochemical parameters (RBC, HGB and WBC) examined before and
after implantation are shown in Figures 11–14. The Zn content was slightly increased in the case of
pure Zn and Zn–1.3%Fe implantations compared to titanium, as expected, but in all cases was within
the normal range of 168–190 [μg/dL] [57]. Furthermore, none of the implanted samples exerted a
negative effect on the hematological profiles of the rats, as all of the measurements were within the
normal range [58]: RBC 7.62–9.99 [106/μL], HGB 13.6–17.4 [g/dL], WBC 1.98–11.06 [103/μL].
The corrosion rate of pure Zn and Zn–1.3%Fe implants at 14 weeks post implantation are shown
in Figure 15. The corrosion rate of Zn–1.3%Fe was significantly increased relative to pure Zn.

165
Metals 2018, 8, 153

Figure 10. Rat body weights pre- and post-implantation.

Figure 11. Serum Zinc level pre- and post-implantation.

Figure 12. Red blood cells (RBC) level pre- and post-implantation.

166
Metals 2018, 8, 153

Figure 13. Hemoglobin (HGB) level pre- and post-implantation.

Figure 14. White blood cells (WBC) level pre- and post-implantation.

Figure 15. Corrosion rate of pure Zn and Zn–1.3%Fe implants after 14 weeks of implantation.

Histology analyses of the subcutaneous tissue around the tested implants (pure Zn, Zn–1.3%Fe
and Ti–6Al–4V as reference alloy) at 14 weeks post-implantation are shown in Figure 16. None of

167
Metals 2018, 8, 153

the zinc-bearing implants promoted a negative effect on the live tissue or provoked an inflammatory
response compared to the reference Ti alloy.

A B

* *
C

*
Figure 16. Histological analysis of subcutaneous tissue in the vicinity of the implants. (A) Pure Zn;
(B) Zn–1.3%Fe; (C) Ti–6Al–4V. * marks the implants location.

4. Discussion
Biodegradable implants that strongly resist corrosion in physiological environments may elicit
negative inflammatory responses, as macrophages increase their activity to metabolize foreign
material [27]. Indeed, earlier in vivo work with zinc vascular implants demonstrated that the
corrosion rate may be a critical regulator of inflammation [27]. In this work, we were surprised
to find that the highest purity zinc implant (~0.01 wt % impurities) produced the strongest negative
inflammatory response, and introductions of mainly iron impurities of ~0.3 wt % into the implant
substantially ameliorated the inflammatory response. This discovery raised the intriguing prospect
that inflammatory responses to zinc implants could be controlled by the implant’s corrosion behavior.
In the present proof of concept work, we intended to develop a zinc alloy system that imparts
tunable corrosion rates (by adjusting alloying element concentration) in order to regulate inflammatory
responses. Herein, we have developed such an implant material as well as characterized the effects
of alloying on material properties. The ultimate goal, not pursued in the present study, is to
identify an optimal corrosion rate for avoiding harmful inflammation when deploying zinc based
medical implants.
Several methods can be used to increase the corrosion rate of pure zinc. These methods
include surface film modification, coatings, microstructure manipulation via thermal and mechanical
processing, and alloying. The intent of the present study was to use the biocompatible alloying element
Fe to produce micro-galvanic effects to increase the corrosion rate of pure zinc. This approach was also
adopted by other researchers, for instance in Zn–Cu base systems such as Zn–Cu, by Niu et al. [40],
Zn–Cu–Mg by Tang et al. [44] and Zn–3Cu by Yue et al. [36]. The use of Cu as an alloying
element to produce a micro-galvanic effect was justified by the higher standard potential of Cu
(0.337 V compared to –0.763 V for Zinc) as well as the micro-galvanic effects generated by a secondary

168
Metals 2018, 8, 153

phase CuZn5 . The potential to achieve a micro-galvanic effect in the present study is based on the
elevated electrochemical potential of all Zn–Fe precipitate phases relative to pure zinc. The potentials of
Zn–Fe phases as identified by Lee et al. [53] are as follows (vs. SCE): Delta phase −0.822 V, Zeta phase
−0.940 V, Gamma phase −0.772 V, while pure zinc is −1.03 V.
In-vitro examination of the gravity cast Zn–1.3%Fe alloy demonstrated a nearly 100% increase
in corrosion rate compared to pure zinc. This was supported by the electrochemical analysis in
terms of potentiodynamic polarization and impedance examination as well as the in vivo corrosion
rate measurements at 14 weeks post implantation. The mechanical properties of pure zinc in as
cast conditions obtained by this study are in line with the reported results of Gong et al. [37],
and Li et al. [39]. The increased strength of Zn–1.3%Fe with concomitant reduction in ductility is
attributed to the effect of Fe as an alloying element. Nevertheless, the strength and ductility of
Zn–1.3%Fe alloy in the as cast condition are still relatively low. In order to improve these properties,
additional metal forming processing, such as hot extrusion [59], are required to obtain an adequate
structural material for practical degradable implant applications.
The in vivo study demonstrated that the pure Zn and Zn–1.3%Fe alloy avoided negative systemic
effects on the wellbeing behavior and normal growth of the rats during the 12 weeks of implantation.
Furthermore, the implantation of these two materials did not increase the amount of zinc in blood
beyond the acceptable level. This was supported by the absence of anemia in terms of RBC and
HGB levels. In addition, there were no signs of infection following implantation of either Zn base
system, as there was no increase in the amount of white blood cells (WBC). This was confirmed by
the histology analyses of subcutaneous tissue in the vicinity of the tested Zn base implants. Although
not performed directly at the implant-tissue interface, the histology shows a general absence of
harmful effects at 14 weeks post implantation from the implant or corrosion activity. The corrosion
rate of pure Zn and Zn–1.3%Fe alloy under in vivo conditions was relatively reduced compared to
their corrosion rate under in vitro conditions. This result is in accordance with the observations
of Torne et al. [56], and Witte et al. [60] who have reported that the aggressiveness of the in vivo
physiological environment towards the biodegradable metal implants is relatively lower compared to
the in vitro submersion of metals under PBS solution. Further contributing to the disparity, the two
cylindrical disks implanted in the midline of the back of the rats were exposed to a relatively low blood
flow that practically generates a relatively reduced corrosion attack compared to the in vitro conditions.
Together, the results of the present study indicate that Zn–1.3%Fe alloy can significantly increase
the corrosion rate of pure Zn. Consequently, this alloy can overcome the inherent low degradation rate
of pure Zn and can be considered as a potential structural material candidate for biodegradable Zn
base implants.

5. Conclusions
Alloying of Zn with Fe at a concentration of 1.3 wt % nearly doubled the corrosion rate in in vitro
and in vivo conditions. The relatively higher corrosion of Zn–1.3%Fe is attributed to the microgalvanic
effect generated mainly by the formation of Delta phase (Zn11 Fe). The absence of undesirable systemic
effects, including weight growth, wellbeing activities and hematological characteristics (RBC, HGB
and WBC) of the rats during 14 weeks of implantation, as well as obtaining adequate histology results
in subcutaneous tissue close to the tested implants suggests that the Zn–1.3%Fe alloy can be considered
as a potential material for biodegradable implants.

Acknowledgments: The authors would like to thank Avi Leon for his assistance in the experimental work.
Author Contributions: Eli Aghion and Alon Kafri conceived and designed the experiments; Alon kafri and
Shira Ovadia performed the experiments; Eli Aghion, Alon Kafri, Jeremy Goldman and Jaroslaw Drelich analyzed
the data; Eli Aghion and Alon Kafri wrote the paper.
Conflicts of Interest: The authors declare no conflict of interest.

169
Metals 2018, 8, 153

References
1. Schinhammer, M.; Hänzi, A.C.; Löffler, J.F.; Uggowitzer, P.J. Design strategy for biodegradable Fe-based
alloys for medical applications. Acta Biomater. 2010, 6, 1705–1713. [CrossRef] [PubMed]
2. Peuster, M.; Wohlsein, P.; Brugmann, M.; Ehlerding, M.; Seidler, K.; Fink, C.; Brauer, H.; Fischer, A.;
Hausdorf, G. A novel approach to temporary stenting: Degradable cardiovascular stents produced from
corrodible metal-results 6–18 months after implantation into New Zealand white rabbits. Heart 2001, 86,
563–569. [CrossRef] [PubMed]
3. Feng, Q.; Zhang, D.; Xin, C.; Liu, X.; Lin, W.; Zhang, W.; Chen, S.; Sun, K. Characterization and in vivo
evaluation of a bio-corrodible nitrided iron stent. J. Mater. Sci. Mater. Med. 2013, 24, 713–724. [CrossRef]
[PubMed]
4. Mueller, P.P.; Arnold, S.; Badar, M.; Bormann, D.; Bach, F.W.; Drynda, A.; Meyer-Lindenberg, A.; Hauser, H.;
Peuster, M. Histological and molecular evaluation of iron as degradable medical implant material in a
murine animal model. J. Biomed. Mater. Res. Part A 2012, 100A, 2881–2889. [CrossRef] [PubMed]
5. Liu, B.; Zheng, Y.F. Effects of alloying elements (Mn, Co, Al, W, Sn, B, C and S) on biodegradability and
in vitro biocompatibility of pure iron. Acta Biomater. 2011, 7, 1407–1420. [CrossRef] [PubMed]
6. Pierson, D.; Edick, J.; Tauscher, A.; Pokorney, E.; Bowen, P.; Gelbaugh, J.; Stinson, J.; Getty, H.; Lee, C.H.;
Drelich, J.; et al. A simplified in vivo approach for evaluating the bioabsorbable behavior of candidate stent
materials. J. Biomed. Mater. Res. Part B Appl. Biomater. 2012, 100B, 58–67. [CrossRef] [PubMed]
7. Bowen, P.K.; Drelich, J.; Buxbaum, R.E.; Rajachar, R.M.; Goldman, J. New approaches in evaluating metallic
candidates for bioabsorbable stents. Emerg. Mater. Res. 2012, 1, 237–255. [CrossRef]
8. Schümann, K.; Ettle, T.; Szegner, B.; Elsenhans, B.; Solomons, N.W. On risks and benefits of iron
supplementation recommendations for iron intake revisited. J. Trace Elem. Med. Biol. 2007, 21, 147–168.
[CrossRef] [PubMed]
9. Staiger, M.P.; Pietak, A.M.; Huadmai, J.; Dias, G. Magnesium and its alloys as orthopedic biomaterials:
A review. Biomaterials 2006, 27, 1728–1734. [CrossRef] [PubMed]
10. Chen, D.; He, Y.; Tao, H.; Zhang, Y.; Jiang, Y.; Zhang, X.; Zhang, S. Biocompatibility of magnesium-zinc alloy
in biodegradable orthopedic implants. Int. J. Mol. Med. 2011, 28, 343–348. [PubMed]
11. Dunne, C.F.; Levy, G.K.; Hakimi, O.; Aghion, E.; Twomey, B.; Stanton, K.T. Corrosion behaviour of
biodegradable magnesium alloys with hydroxyapatite coatings. Surf. Coat. Technol. 2016, 289, 37–44.
[CrossRef]
12. Hakimi, O.; Aghion, E. Corrosion performance of biodegradable Mg-6%Nd-2%Y-0.5%Zr produced by melt
spinning technology. Adv. Eng. Mater. 2014, 16, 364–370. [CrossRef]
13. Zhang, S.; Zhang, X.; Zhao, C.; Li, J.; Song, Y.; Xie, C.; Tao, H.; Zhang, Y.; He, Y.; Jiang, Y. Research on an
Mg-Zn alloy as a degradable biomaterial. Acta Biomater. 2010, 6, 626–640. [CrossRef] [PubMed]
14. Persaud-Sharma, D.; McGoron, A. Biodegradable Magnesium Alloys: A Review of Material Development
and Applications. J. Biomim. Biomater. Tissue Eng. 2012, 12, 25–39. [CrossRef] [PubMed]
15. Song, G. Control of biodegradation of biocompatable magnesium alloys. Corros. Sci. 2007, 49, 1696–1701.
[CrossRef]
16. Ghali, E. Testing of Aluminum, Magnesium, and Their Alloys. In Uhlig’s Corrosion Handbook, 3rd ed.;
John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2011; pp. 1103–1106.
17. Hakimi, O.; Aghion, E.; Goldman, J. Improved stress corrosion cracking resistance of a novel biodegradable
EW62 magnesium alloy by rapid solidification, in simulated electrolytes. Mater. Sci. Eng. C 2015, 51, 226–232.
[CrossRef] [PubMed]
18. Aghion, E.; Levy, G. The effect of Ca on the in vitro corrosion performance of biodegradable Mg-Nd-Y-Zr
alloy. J. Mater. Sci. 2010, 45, 3096–3101. [CrossRef]
19. Levy, G.; Aghion, E. Effect of diffusion coating of Nd on the corrosion resistance of biodegradable Mg
implants in simulated physiological electrolyte. Acta Biomater. 2013, 9, 8624–8630. [CrossRef] [PubMed]
20. Cheng, J.; Liu, B.; Wu, Y.H.; Zheng, Y.F. Comparative invitro study on pure metals (Fe, Mn, Mg, Zn and W)
as biodegradable metals. J. Mater. Sci. Technol. 2013, 29, 619–627. [CrossRef]
21. Bowen, P.K.; Shearier, E.R.; Zhao, S.; Guillory, R.J., II; Zhao, F.; Goldman, J.; Drelich, J.W. Biodegradable
Metals for Cardiovascular Stents: From Clinical Concerns to Recent Zn-Alloys. Adv. Healthc. Mater. 2016, 5,
1121–1140. [CrossRef] [PubMed]

170
Metals 2018, 8, 153

22. Bowen, P.K.; Drelich, J.; Goldman, J. Zinc exhibits ideal physiological corrosion behavior for bioabsorbable
stents. Adv. Mater. 2013, 25, 2577–2582. [CrossRef] [PubMed]
23. Hennig, B.; Toborek, M.; McClain, C.J. Antiatherogenic properties of zinc: Implications in endothelial cell
metabolism. Nutrition 1996, 12, 711–717. [CrossRef]
24. Al-Maroof, R.A.; Al-Sharbatti, S.S. Serum zinc levels in diabetic patients and effect of zinc supplementation
on glycemic control of type 2 diabetics. Saudi Med. J. 2006, 27, 344–350. [PubMed]
25. Hambidge, K.M.; Krebs, N.F. Zinc Deficiency: A Special Challenge. J. Nutr. 2007, 137, 1101–1105. [CrossRef]
[PubMed]
26. Plum, L.M.; Rink, L.; Hajo, H. The essential toxin: Impact of zinc on human health. Int. J. Environ. Res.
Public Health 2010, 7, 1342–1365. [CrossRef] [PubMed]
27. Guillory, R.J.; Bowen, P.K.; Hopkins, S.P.; Shearier, E.R.; Earley, E.J.; Gillette, A.A.; Aghion, E.; Bocks, M.;
Drelich, J.W.; Goldman, J. Corrosion Characteristics Dictate the Long-Term Inflammatory Profile of
Degradable Zinc Arterial Implants. ACS Biomater. Sci. Eng. 2016, 2, 2355–2364. [CrossRef]
28. Drelich, A.J.; Zhao, S.; Guillory, R.J.; Drelich, J.W.; Goldman, J. Long-term surveillance of zinc implant in
murine artery: Surprisingly steady biocorrosion rate. Acta Biomater. 2017, 58, 539–549. [CrossRef] [PubMed]
29. Bowen, P.K.; Guillory, R.J., II; Shearier, E.R.; Seitz, J.-M.; Drelich, J.; Bocks, M.; Zhao, F.; Goldman, J. Metallic
zinc exhibits optimal biocompatibility for bioabsorbable endovascular stents. Mater. Sci. Eng. C 2015, 56,
467–472. [CrossRef] [PubMed]
30. Smith, S.E.; Lakson, E.J. Zinc toxicity in rats; antagonistic effects of copper and liver. J. Biol. Chem. 1946, 163,
29–38. [PubMed]
31. Piao, F.; Yokoyama, K.; Ma, N.; Yamauchi, T. Subacute toxic effects of zinc on various tissues and organs of
rats. Toxicol. Lett. 2003, 145, 28–35. [CrossRef]
32. Drelich, A.J.; Bowen, P.K.; LaLonde, L.; Goldman, J.; Drelich, J.W. Importance of oxide film in endovascular
biodegradable zinc stents. Surf. Innov. 2016, 4, 133–140. [CrossRef]
33. Shomali, A.A.; Guillory, R.J.; Seguin, D.; Goldman, J.; Drelich, J.W. Effect of PLLA coating on corrosion and
biocompatibility of zinc in vascular environment. Surf. Innov. 2017, 5, 211–220. [CrossRef]
34. Tjellström, A.; Rosenhall, U.; Lindström, J.; Hallén, O.; Albrektsson, T.; Brånemark, P.I. Five-Year Experience
with Skin-Penetrating Bone-Anchored Implants in the Temporal Bone. Acta Otolaryngol. 1983, 95, 568–575.
[CrossRef] [PubMed]
35. Geetha, M.; Singh, A.K.; Asokamani, R.; Gogia, A.K. Ti based biomaterials, the ultimate choice for orthopaedic
implants—A review. Prog. Mater. Sci. 2009, 54, 397–425. [CrossRef]
36. Yue, R.; Huang, H.; Ke, G.; Zhang, H.; Pei, J.; Xue, G.; Yuan, G. Microstructure, mechanical properties and
in vitro degradation behavior of novel Zn-Cu-Fe alloys. Mater. Charact. 2017, 134, 114–122. [CrossRef]
37. Gong, H.; Wang, K.; Strich, R.; Zhou, J.G. In vitro biodegradation behavior, mechanical properties,
and cytotoxicity of biodegradable Zn-Mg alloy. J. Biomed. Mater. Res. Part B Appl. Biomater. 2015, 103,
1632–1640. [CrossRef] [PubMed]
38. Vojtěch, D.; Kubásek, J.; Šerák, J.; Novák, P. Mechanical and corrosion properties of newly developed
biodegradable Zn-based alloys for bone fixation. Acta Biomater. 2011, 7, 3515–3522. [CrossRef] [PubMed]
39. Li, H.F.; Xie, X.H.; Zheng, Y.F.; Cong, Y.; Zhou, F.Y.; Qiu, K.J.; Wang, X.; Chen, S.H.; Huang, L.; Tian, L.; et al.
Development of biodegradable Zn-1X binary alloys with nutrient alloying elements Mg, Ca and Sr. Sci. Rep.
2015, 5, 10719. [CrossRef] [PubMed]
40. Zhao, S.; McNamara, C.T.; Bowen, P.K.; Verhun, N.; Braykovich, J.P.; Goldman, J.; Drelich, J.W. Structural
Characteristics and In Vitro Biodegradation of a Novel Zn-Li Alloy Prepared by Induction Melting and Hot
Rolling. Metall. Mater. Trans. A Phys. Metall. Mater. Sci. 2017, 48, 1204–1215. [CrossRef]
41. Bowen, P.K.; Seitz, J.M.; Guillory, R.J., II; Braykovich, J.P.; Zhao, S.; Goldman, J.; Drelich, J.W. Evaluation of
wrought Zn–Al alloys (1, 3, and 5 wt % Al) through mechanical and in vivo testing for stent applications.
J. Biomed. Mater. Res. Part B Appl. Biomater. 2018, 106, 245–258. [CrossRef] [PubMed]
42. Zhao, S.; Seitz, J.M.; Eifler, R.; Maier, H.J.; Guillory, R.J., II; Earley, E.J.; Drelich, A.; Goldman, J.; Drelich, J.W.
Zn-Li alloy after extrusion and drawing: Structural, mechanical characterization, and biodegradation in
abdominal aorta of rat. Mater. Sci. Eng. C 2017, 76, 301–312. [CrossRef] [PubMed]
43. Niu, J.; Tang, Z.; Huang, H.; Pei, J.; Zhang, H.; Yuan, G.; Ding, W. Research on a Zn-Cu alloy as a
biodegradable material for potential vascular stents application. Mater. Sci. Eng. C 2016, 69, 407–413.
[CrossRef] [PubMed]

171
Metals 2018, 8, 153

44. Tang, Z.; Huang, H.; Niu, J.; Zhang, L.; Zhang, H.; Pei, J.; Tan, J.; Yuan, G. Design and characterizations of
novel biodegradable Zn-Cu-Mg alloys for potential biodegradable implants. Mater. Des. 2017, 117, 84–94.
[CrossRef]
45. Bakhsheshi-Rad, H.R.; Hamzah, E.; Low, H.T.; Kasiri-Asgarani, M.; Farahany, S.; Akbari, E.; Cho, M.H.
Fabrication of biodegradable Zn-Al-Mg alloy: Mechanical properties, corrosion behavior, cytotoxicity and
antibacterial activities. Mater. Sci. Eng. C 2017, 73, 215–219. [CrossRef] [PubMed]
46. Jin, H.; Zhao, S.; Guillory, R.; Bowen, P.K.; Yin, Z.; Griebel, A.; Schaffer, J.; Earley, E.J.; Goldman, J.; Drelich, J.W.
Novel high-strength, low-alloys Zn-Mg (<0.1 wt % Mg) and their arterial biodegradation. Mater. Sci. Eng. C
2017, 84, 67–79.
47. Dambatta, M.S.; Izman, S.; Kurniawan, D.; Hermawan, H. Processing of Zn-3Mg alloy by equal channel
angular pressing for biodegradable metal implants. J. King Saud Univ. Sci. 2017, 29, 455–461. [CrossRef]
48. Mostaed, E.; Sikora-Jasinska, M.; Mostaed, A.; Loffredo, S.; Demir, A.G.; Previtali, B.; Mantovani, D.;
Beanland, R.; Vedani, M. Novel Zn-based alloys for biodegradable stent applications: Design, development
and in vitro degradation. J. Mech. Behav. Biomed. Mater. 2016, 60, 581–602. [CrossRef] [PubMed]
49. Bagha, P.S.; Khaleghpanah, S.; Sheibani, S.; Khakbiz, M.; Zakeri, A. Characterization of nanostructured
biodegradable Zn-Mn alloy synthesized by mechanical alloying. J. Alloys Compd. 2018, 735, 1319–1327.
[CrossRef]
50. Xiao, C.; Wang, L.; Ren, Y.; Sun, S.; Zhang, E.; Yan, C.; Liu, Q.; Sun, X.; Shou, F.; Duan, J.; et al. Indirectly
extruded biodegradable Zn-0.05wt%Mg alloy with improved strength and ductility: In vitro and in vivo
studies. J. Mater. Sci. Technol. 2018. [CrossRef]
51. Zhang, Y.; Li, J.; Li, J. Microstructure, mechanical properties, corrosion behavior and film formation
mechanism of Mg-Zn-Mn-xNd in Kokubo’s solution. J. Alloys Compd. 2018, 730, 458–470. [CrossRef]
52. Fransen, M.; Nazikkol, C. Zinc/Iron Phase Transformation Studies on Galvannealed Steel Coatings By X-ray
Diffraction. Advances 2003, 46, 291–296.
53. Lee, H.H.; Hiam, D. Corrosion resistance of galvannealed steel. Corrosion 1989, 45, 852–856. [CrossRef]
54. Yun, Y.; Dong, Z.; Yang, D.; Schulz, M.J.; Shanov, V.N.; Yarmolenko, S.; Xu, Z.; Kumta, P.; Sfeir, C.
Biodegradable Mg corrosion and osteoblast cell culture studies. Mater. Sci. Eng. C 2009, 29, 1814–1821.
[CrossRef]
55. Jaiswal, S.; Kumar, R.M.; Gupta, P.; Kumaraswamy, M.; Roy, P.; Lahiri, D. Mechanical, corrosion and
biocompatibility behaviour of Mg-3Zn-HA biodegradable composites for orthopaedic fixture accessories.
J. Mech. Behav. Biomed. Mater. 2018, 78, 442–454. [CrossRef] [PubMed]
56. Törne, K.; Larsson, M.; Norlin, A.; Weissenrieder, J. Degradation of zinc in saline solutions, plasma,
and whole blood. J. Biomed. Mater. Res. Part B Appl. Biomater. 2016, 104, 1141–1151. [CrossRef] [PubMed]
57. Yur, F.; Bildik, A.; Belge, F.; Kilicalp, D. Serum Plasma and Erythrocyte Zinc levels in various animal species.
VAN Vet. J. 2002, 13, 82–83.
58. Quesenberry, K.E.; Carpenter, J.W. Small Rodents. In Ferrets, Rabbits and Rodents: Clinical Medicine and Surgery;
Elsevier: St. Louis, MO, USA, 2003; p. 348.
59. Sikora-Jasinska, M.; Mostaed, E.; Mostaed, A.; Beanland, R.; Mantovani, D.; Vedani, M. Fabrication,
mechanical properties and in vitro degradation behavior of newly developed Zn Ag alloys for degradable
implant applications. Mater. Sci. Eng. C 2017, 77, 1170–1181. [CrossRef] [PubMed]
60. Witte, F.; Fischer, J.; Nellesen, J.; Crostack, H.A.; Kaese, V.; Pisch, A.; Beckmann, F.; Windhagen, H. In vitro
and in vivo corrosion measurements of magnesium alloys. Biomaterials 2006, 27, 1013–1018. [CrossRef]
[PubMed]

© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

172
metals
Article
Preparation and Characterization of Zinc Materials
Prepared by Powder Metallurgy
Michaela Krystýnová 1, *, Pavel Doležal 1,2 , Stanislava Fintová 1,3 , Matěj Březina 1 , Josef Zapletal 2
and Jaromír Wasserbauer 1
1 Materials Research Centre, Faculty of Chemistry, Brno University of Technology, Purkyňova 464/118,
612 00 Brno, Czech Republic; dolezal@fme.vutbr.cz (P.D.); fintova@ipm.cz (S.F.);
xcbrezinam@fch.vut.cz (M.B.); wasserbauer@fch.vut.cz (J.W.)
2 Institute of Materials Science and Engineering, Faculty of Mechanical Engineering,
Brno University of Technology, Technická 2896/2, 616 69 Brno, Czech Republic; zapletal@fme.vutbr.cz
3 Institute of Physics of Materials Academy of Sciences of the Czech Republic, Žižkova 22,
616 62 Brno, Czech Republic
* Correspondence: xckrystynovam@fch.vut.cz; Tel.: +420-541-149-469

Received: 30 August 2017; Accepted: 22 September 2017; Published: 27 September 2017

Abstract: The use of zinc-based materials as biodegradable materials for medical purposes is offered
as a possible alternative to corrosion-less resistant magnesium-based materials. Zinc powders with
two different particle sizes (7.5 μm and 150 μm) were processed by the methods of powder metallurgy:
cold pressing, cold pressing followed by sintering and hot pressing. The microstructure of prepared
materials was evaluated in terms of light optical microscopy, and the mechanical properties were
analyzed with Vickers microhardness testing and three-point bend testing. Fractographic analysis
of broken samples was performed with scanning electron microscopy. Particle size was shown to
have a significant effect on compacts mechanical properties. The deformability of 7.5 μm particle
size powder was improved by increased temperature during the processing, while in the case of
larger powder, no significant influence of temperature was observed. Bending properties of prepared
materials were positively influenced by elevated temperature during processing and correspond to
the increasing compacting pressures. Better properties were achieved for pure zinc prepared from
150 μm particle size powder compared to materials prepared from 7.5 μm particle size powder.

Keywords: powder metallurgy; zinc; cold pressing; sintering; hot pressing; mechanical properties

1. Introduction
In the field of biodegradable metal materials, magnesium and its alloys are the most studied [1–6].
Due to their suitable mechanical properties such as high specific strength, stiffness and damping
ability, they are suitable materials for the preparation of bone implants [1–7]. Because of its low
corrosion resistance and therefore difficult degradation control, magnesium is not used in its pure
state, but alloyed. Commercial magnesium alloys are not primarily designed for medical applications,
but some studies on the biomedical purpose of these alloys were done before, showing good corrosion
behavior and biocompatibility [8,9]. For medical purposes, magnesium alloys with calcium, zinc, rare
earth elements or manganese are being considered [10–13]. Due to the low density of calcium, these
alloys have a similar density as a natural bone. In addition, Ca2+ ions are beneficial for human bones,
and Mg2+ ions support the function of Ca2+ ions and generally the treatment of injury [11,14,15].
Zinc has been considered, in the field of biomedical applications, a suitable alloying element
for magnesium alloys in terms of improving corrosion resistance and enhancement of mechanical
properties [10,16,17]. Considering zinc is a nobler metal than magnesium and zinc is also biocompatible,
it can be a suitable biodegradable material [16–18]. Zinc is a nutritionally-essential element in the

Metals 2017, 7, 396; doi:10.3390/met7100396 173 www.mdpi.com/journal/metals


Metals 2017, 7, 396

human body; approximately 85% of Zn in the human body can be found in bone and muscle. It is an
integral part of the structure of macromolecules and enzymes, and it participates in a large number
of enzymatic reactions. The human zinc requirement for adult males is 10–15 mg/day (upper limit
40 mg/day) [19,20]. The potential for systemic toxicity of metallic zinc should be nonexistent due to
the rapid transport of ionic zinc in living tissue. Moreover, higher consumption (up to 100 mg/day)
of zinc is considered non-toxic. However, more detailed cytotoxicity tests must be carried out in the
future [15,21–23].
Based on its positive effect on the human body, zinc can be considered a suitable base material for
the preparation of biodegradable materials. There are only a few works dealing with the use of zinc
alloys for biomedical applications. In the case of pure zinc, there is just one article [22] dealing with
the applicability of zinc as a bioabsorbable cardiac stent material, according to the author’s knowledge.
The authors in [22] demonstrated the biocorrosion behavior of zinc by a series of four wire samples
implanted in the abdominal aorta of Sprague-Dawley rat for 1.5, 3, 4.5 and 6 months. Results show
that zinc corrosion is slower than the corrosion of magnesium and its alloys [24,25]. Zinc implants stay
almost intact for approximately four months, and then, corrosion subsequently accelerates. Corrosion
products of zinc are formed from zinc oxide and zinc carbonate. Due to the slow corrosion of zinc,
there is sufficient time to exclude hydrogen resulting from the corrosion process. That is both the
main difference and an advantage over magnesium [21]. However, no mechanical tests results were
provided in this work.
Generally, the mechanical properties of zinc are relatively low. The modulus of compressive
elasticity of commercially pure zinc is not listed as an exact value because there is no region of strict
proportionality in the compressive stress-strain curve. Therefore, this value was determined in the
range from 70–140 GPa [25]. The Vickers hardness of this material is 30 HV. The tensile strength
of wrought pure zinc is in the range of 120–150 MPa, depending on the direction of rolling [26].
The tensile strength of cast pure zinc is very low with the value approximately 25 MPa [17]. Another
article dealing with mechanical properties of zinc showed that applying hot extrusion (300 ◦ C, extrusion
ratio 10:1, 2 mm/min) can enhance the ultimate tensile strength to a value of approximately 100 MPa
and a Vickers hardness to 44 HV5 [27] and, in [28], to approximately 110 MPa for hot extruded pure zinc
material (250 ◦ C, 14:1). Due to the poor mechanical properties, alloying of zinc is a suitable solution.
As alloying elements, mostly magnesium [17,27,29], aluminum [30,31] and silver [28] were studied.
Alloying zinc with aluminum provides substantial enhancement of the mechanical properties.
That is due to the presence of Al-Zn solid solution and the volume fraction of the lamellar
microconstituents from the monotectoid reaction acting as potential barriers for dislocation motion.
With an addition of 5.5 wt % of aluminum, the tensile strength of material prepared by hot rolling
at 350 ◦ C reaches the value of approximately 308 MPa and a yield strength of approximately
240 MPa [30]. However, the use of Zn-Al alloys in medicine is still limited, because of the uncertainty
regarding the toxicity of aluminum. Although, the toxicity of aluminum has never been sufficiently
proven [30], its connection to neurological disorders is still discussed in the literature [32]. Furthermore,
the corrosion resistance of Zn-Al alloys tended to be lower compared to high purity zinc due to the
intergranular corrosion. Moreover, volume expansion associated with the formation of corrosion
products led to cracking and fragmentation of implants [30]. Another alloying element with positive
influence on mechanical properties and preserving biocompatibility is silver. With the addition of
7.0 wt % of Ag, the ultimate tensile strength of the cast zinc alloy increased up to 287 MPa [28].
Moreover, silver has been used in medicine for healing wounds, and it is used, in the form of
nanoparticles, for the prevention of the adherence of bacteria to the surface of implants [33]. However,
the presence of secondary phase particles in the Zn-Ag alloys led to micro-galvanic corrosion because
they act as anodes. Consequently, Zn-Ag alloys have higher degradation rates in comparison with
pure Zn [28].
Probably the most discussed alloying element of zinc-based biodegradable materials is
magnesium. Because of the existence of the hard Mg2 Zn11 intermetallic phase in the alloy structure,

174
Metals 2017, 7, 396

the value of the hardness increases with a higher content of magnesium up to the value of 200 HV (cast
alloy, 3 wt % of magnesium). With higher magnesium content (35–45 wt % of magnesium), the hardness
of the cast alloys can reach values of approximately 285–300 HV1 for 35 wt % of magnesium (depending
on the cast product cooling rate and subsequent microstructure). The hardness increases because of
the presence of another strengthening phase in the alloy microstructure (MgZn2 ). With the higher
content of magnesium in the alloy, the content of the MgZn2 phase decreases and the content of the
MgZn and Mg7 Zn3 phases increases. The hardness falls again due to decreasing content of MgZn2 to
255–280 HV1 for the alloy with 45 wt % of magnesium (depending on the cast product cooling rate and
subsequent microstructure) [29]. However, the existence of brittle eutectic phases in alloys with the
content of magnesium higher than 1 wt % has a negative effect on ultimate tensile strength. Ultimate
tensile strength increases up to 150 MPa (1 wt % of Mg) and then decreases to the value of 30 MPa
(3 wt % of Mg). That is the same value as for pure zinc prepared by the same method. Furthermore,
elongation of Zn-Mg alloys reaches the highest value for the alloy with 1 wt % of magnesium [17].
Another article deals with an enhancement of mechanical properties with hot extrusion processing.
Hot extruded alloy with 0.8 wt % of magnesium reaches the ultimate tensile strength of approximately
300 MPa and a Vickers hardness of approximately 80 HV5. With the content of 1.6 wt % of magnesium,
hot extruded materials reach the ultimate tensile strength of approximately 360 MPa and a Vickers
hardness of approximately 97 HV5 [27].
Most of the available articles deal with the influence of alloying elements and their content on
the mechanical properties and corrosion resistance of zinc-based materials prepared by casting or
mechanical treatment of cast products. The influence of the preparation method and its parameters on
mechanical properties is studied only marginally [17,27–30].
This work deals with three different methods of preparation of zinc materials by powder
metallurgy: cold pressing, cold pressing followed by sintering and hot pressing. According to the
author’s knowledge, there is no available publication related to the preparation of pure zinc by these
methods. Besides the influence of the preparation method, this work also focuses on the influence of the
particle size of the used powder materials on the resulting microstructure and mechanical properties.

2. Materials and Methods


For the experiment, two different zinc powders (99.8% purity, 7.5 μm and 150 μm mean particle
size provided by Goodfellow Cambridge Limited Company (Huntingdon, UK) were used. The smaller,
7.5 μm particle size powder (Zn7.5) was prepared by the electrothermal process, and the larger, 150 μm
particle size powder (Zn150) was prepared by air atomization. In order to prevent material oxidation,
manipulation with Zn powder was carried out in an inert atmosphere (N2 ) in the glove box. The size
and shape of metal powder particles were verified and analyzed by a Zeiss Evo LS 10 scanning electron
microscope (SEM; Carl Zeiss Ltd., Cambridge, UK).
The particles of Zn powder with the declared particle size of 7.5 μm had spherical shapes, and their
size ranged from 1 μm to a maximum value of 20 μm (Figure 1a). The particles tended to clump. The
smaller particles attached to the larger ones and created clusters. The particles of zinc powder with a
declared particle size of 150 μm were irregularly rod-shaped with a minor amount of round-shaped
particles (Figure 1b). The SEM analysis of the powder showed that some particles in their largest
dimension reached a size from 40–640 μm.
The purity of powders declared by the supplier was verified by EDS analysis. The larger content
of oxygen, approximately 8.0 ± 0.5 wt %, was determined in Zn7.5 powder when compared to Zn150
powder, containing only 2.5 ± 0.5 wt %.
Experimental samples were processed in three ways: (i) By bidirectional cold pressing (CP),
(ii) cold pressing followed by sintering (CP-S) and (iii) hot pressing (HP). A hollow cylindrical steel
die with the inner diameter of 20 mm was used for zinc powders’ compaction. Before it was filled
with zinc, the die surface was carburized in order to prevent adhesion of the base powder material to
the surface of the die. The pressing of the powders was carried out by the Zwick Z250 Allround-Line

175
Metals 2017, 7, 396

universal testing machine (Zwick GmbH & Co.KG, Ulm, Germany) with a velocity of 2 mm/min.
Pressures of 100, 200, 300, 400 and 500 MPa were used for prepared powders’ compaction.

(a) (b)

Figure 1. Zn particles, SEM (scanning electron microscope): (a) Mean particle size 7.5 μm; (b) mean
particle size 150 μm.

For sintering, cold-pressed samples pressed under 100, 200, 300, 400 and 500 MPa were used.
Cold-pressed samples were inserted into glass vials and then filled with argon of purity 4.6 and sealed.
The sintering of pressed powder was done for 1 h at a temperature of 400 ◦ C in the laboratory furnace
preheated to the required temperature.
Hot pressing was carried out for 1 h at a temperature of 400 ◦ C under 100, 200, 300, 400 and
500 MPa. Prepared bulk materials were in the form of cylindrical tablets with a diameter of 20 mm
and a height of 5 mm.
Metallographic evaluation of prepared samples was performed in a conventional manner.
Isopropanol was used as a lubricant and a rinse to prevent oxidation of the samples during grinding and
polishing. For metallographic evaluation, the Zeiss Axio Z1M (Carl Zeiss AG, Oberkochen, Germany)
inverted light optical microscope (OM) and Zeiss Evo LS 10 (Oxford Instruments, Abington, UK) were
used. The microhardness of prepared samples was measured with the LM 248 at machine produced
by LECO Company (Saint Joseph, MO, USA). The measurement was carried out in accordance with
the ISO 6507-1 standard (Vickers method, applied load 25 g) on 10 positions on the sample.
The 3-point bend test was carried out on the Zwick Z020 (Zwick GmbH & Co.KG, Ulm, Germany)
universal testing machine according to the ISO 7438 standard. One sample for each preparation
condition was used for the 3-point bend test. Samples for testing were detracted from the central part
of prepared tablets and ground up to 4 mm × 4 mm × 18 mm proportions with the support span of
16 mm. Fractographic evaluation of the fracture surfaces of broken samples was performed by SEM.
Documentation was carried out in the area of applied tensile stress.

3. Results

3.1. Processed Material Analysis


Microstructures of materials prepared from Zn7.5 applying minimal and maximal (100 and
500 MPa) compacting pressures are summarized in Figure 2. The results showed the low deformation
ability of the Zn7.5 powder particles and an open porosity of materials prepared under the low
compacting pressure of 100 MPa. The porosity of the samples compacted under 100 MPa decreased
with applying temperature. While in the case of CP, quite a high level of porosity was observed, in the
case of the CP-S sample, the porosity level decreased, and only very limited porosity was observed
for the HP Zn7.5 powder sample. Low compacting pressure resulted also in low handling strength
of samples, and falling off of individual particles during the metallographic sample preparation was

176
Metals 2017, 7, 396

observed (marked by arrows). The HP Zn7.5 powder sample microstructural analysis revealed small
particle deformation during the bulk material processing, which was not observed in the case of CP
and CP-S samples compacted under 100 MPa.
The microstructure of materials pressed under 500 MPa shows much lower porosity than samples
pressed under 100 MPa, considering the same processing conditions (Figure 2). Only small particles’
deformation can be seen in the case of the CP sample compacted under 500 MPa (Figure 2b); however,
by the addition of temperature (CP-S and HP), the particles’ deformation increased (Figure 2e,f).

(a) (b)

(c) (d)

(e) (f)

Figure 2. Microstructure of materials prepared from Zn7.5 powder, etched with 5% Nital, optical
microscope (OM): (a) 100 MPa, CP; (b) 500 MPa, CP; (c) 100 MPa, CP-S; (d) 500 MPa, CP-S; (e) 100 MPa,
HP; (f) 500 MPa, HP (CP: Cold pressing; S: Sintering; HP: Hot pressing).

The microstructure of materials prepared from Zn150 powder under 100 and 500 MPa compacting
pressures is shown in Figure 3. In the case of material compacted under a pressure of 100 MPa, closed
pores can be seen between the powder particles in the microstructure (Figure 3a). Even larger porosity

177
Metals 2017, 7, 396

can be seen in the CP-S sample’s microstructure (Figure 3c), while the porosity is not visible in the
case of the HP sample (Figure 3e). Compacting pressure of 500 MPa resulted in a microstructure
without defects for all used processing methods of Zn150 powder-based materials (Figure 3). Zn150
powder particles’ deformation was observed for all the material processing methods and all the used
compacting pressures. Particles deformation increased with the addition of temperature, i.e., larger
particles’ deformation was observed for material processed by CP-S and even larger for material
processed by HP when compared to CP material. Larger particle deformation was observed for
samples prepared under 500 MPa compared to samples prepared under 100 MPa.

(a) (b)

(c) (d)

(e) (f)

Figure 3. Microstructure of materials prepared from Zn150 powder, etched with 5% Nital, OM: (a) 100 MPa,
CP; (b) 500 MPa, CP; (c) 100 MPa, CP-S; (d) 500 MPa, CP-S; (e) 100 MPa, HP; (f) 500 MPa, HP.

178
Metals 2017, 7, 396

The values of microhardness of prepared materials were in the range from 40–49 HV025 for Zn7.5
powder and from 40–45 HV025 for Zn150 powder. No significant dependence on compacting pressure
or processing method was found.

3.2. 3-Point Bend Test


The plots representing the evolution of the flexural strength and the maximal displacement before
the failure are shown in Figure 4. In some cases, no data are plotted due to the low handling strength
of prepared samples or very low obtained values (impossible to determine).
In the case of materials prepared from Zn7.5 powder, there is a clear increasing dependence of
flexural strength on the applied compacting pressure. The same tendency of materials prepared from
Zn150 powder processed by CP and CP-S can be observed. The different behavior appears only in
the case of Zn150 powder HP material where the value of the flexural strength of material processed
under 100 and 200 MPa is around 200 MPa, and then, the value decreased to approximately 150 MPa
for compacting pressures of 300 and 400 MPa. Material pressed under 500 MPa reaches the value of
about 200 MPa again. Generally, CP and CP-S materials prepared from Zn150 powder have higher
flexural strength values than materials from Zn7.5 powder prepared with the same method at the
same compacting pressures. In the case of materials prepared under compacting pressure of 500 MPa
from Zn7.5 powder, the flexural strength increases from 20 MPa for CP to 76 MPa for CP-S material.
The flexural strength of materials prepared from Zn150 powder under the same condition increases
from 88 MPa for CP to 148 MPa for CP-S material.
The values of the flexural strength of materials prepared by HP are substantially higher when
compared to the CP-S and CP materials, especially in the case of materials prepared from Zn7.5
powder. The highest value of flexural strength (322 MPa) is reached with material prepared by HP
under compacting pressure of 400 MPa. While in the case of material from Zn7.5 powder, a significant
increase of flexural strength is observed for all the applied compacting pressures due to the HP
processing (Figure 4a), in the case of Zn150 powder material, no such significant increase of flexural
strength is observed, and in the case of compacting pressure of 400 MPa, the values of flexural strength
for CP-S and HP material are similar (Figure 4b).

(a) (b)

(c) (d)

Figure 4. Dependence of flexural strength and displacement on compacting pressure and processing
of prepared materials: (a) Flexural strength, Zn7.5 powder; (b) Flexural strength, Zn150 powder;
(c) Displacement, Zn7.5 powder; (d) Displacement, Zn150 powder.

179
Metals 2017, 7, 396

The displacement before the failure of Zn7.5 powder materials prepared by CP and by CP-S is
very low (from 0.04–0.10 mm for 500 MPa compacting pressure) (Figure 4c). However, in the case of
HP Zn7.5 powder materials, the displacement before the failure lies in the range from 0.11 mm for the
lowest compacting pressure up to 0.92 mm for the pressure of 500 MPa.
The results of maximal displacement before the failure of materials prepared from Zn150 powder
(Figure 4d) are relatively steady around the value of 0.2 mm for CP and CP-S materials and around
0.4 mm for HP materials. In the case of HP Zn150 powder material, the values of displacement before
failure decrease from 0.49 mm to 0.29 mm for materials compacted under 200 MPa and 400 MPa,
respectively (equivalent to the change of flexural strength) and then again increases for material
pressed under 500 MPa (Figure 4d).

3.3. Fractographic Analysis


The samples broken during the three-point bend testing were examined with the aim to analyze
the fracture mechanism and its correlation with the samples’ processing in terms of SEM. Details
of fracture surfaces of materials prepared from Zn7.5 powder are shown in Figure 5. Only samples
prepared under the lowest and the highest compacting pressures (as possible) with adequate processing
were analyzed. The fracture surface of CP material is characterized by crack growth along powder
particle boundaries, which is also the case for the material compacted under 200 MPa and sintered
at 400 ◦ C (Figure 5a–c). The fracture mechanism can be considered as transgranular. There is visible
porosity between individual powder particles in the case of CP materials and CP-S materials compacted
under 200 MPa. The fracture surface of material pressed under 500 MPa and then sintered shows
lower porosity compared to the previous material states. In addition, crack growth through particle
boundaries is accompanied by powder particle cracking (Figure 5d). The combination of transgranular
with more pronounced intergranular fracture is characteristic for HP materials. Broken powder
particles are characterized by plane facet or by fine non-well developed cleavage facets with a river-like
morphology. HP resulted also in the decrease of a number of pores present on the samples’ fracture
surface, while no visible pores are present on the fracture surface of HP material compacted under
500 MPa (Figure 5f) (showing also the detail of the fracture mechanism).

(a) (b)

(c) (d)

Figure 5. Cont.

180
Metals 2017, 7, 396

(e) (f)

Figure 5. Microstructure of materials prepared from Zn7.5 powder, SEM: (a) 400 MPa, CP; (b) 500 MPa,
CP; (c) 200 MPa, CP-S; (d) 500 MPa, CP-S; (e) 100 MPa, HP; (f) 500 MPa, HP.

In Figure 6, a summary of the details of the fracture surfaces of materials prepared from Zn150
powder is shown. Compared with materials prepared from Zn7.5 powder, in the case of Zn150 powder
materials, there is a visibly higher influence of compacting pressure on particles’ deformation and
porosity. In the case of CP materials, the particles are deformed and mechanically bonded (Figure 6a,b).
Three-point bend test loading led to particles being pulled out of the structure, while holes remaining
on the fracture surface can be seen instead of missing particles in Figure 6a. Similarly to the Zn7.5
powder materials, the transgranular fracture mechanism is characteristic for CP and CP-S samples
compacted under 200 MPa (Figure 6a–c). The combination of transgranular with intergranular fracture
is characteristic for CP-S compacted under 500 MPa and HP materials (Figure 6d–f). Well-developed
cleavage facets with the river-like morphology can be observed on fracture surfaces (broken powder
particles) of HP materials (Figure 6e,f), showing also the detail of the fracture mechanism in Figure 6f.

(a) (b)

(c) (d)

Figure 6. Cont.

181
Metals 2017, 7, 396

(e) (f)

Figure 6. Microstructure of materials prepared from Zn150 powder, SEM: (a) 200 MPa, CP; (b) 500 MPa,
CP; (c) 100 MPa, CP-S; (d) 500 MPa, CP-S; (e) 100 MPa, HP; (f) 500 MPa, HP.

4. Discussion
Microstructural analysis of materials prepared from Zn7.5 powder (Figure 2) revealed lower
deformability of powder particles when compared to Zn150 powder particles (Figure 3). The considerably
lower particle size and spherical shape of Zn7.5 powder may be the reason for this behavior.
Even though the mechanical properties of pure zinc powder particles are comparable, the size
of the particles plays a similar role as in the case of grain size. The larger size of the particles was
connected with their higher deformability during the processing, and the lower particle size resulted
in grain boundaries’ (particles boundaries) strengthening of material. The strengthening of material
was also connected with lower deformability of the material.
The deformability of individual powder particles influenced the material properties and content
of porosity. Although the zinc particles were not significantly deformed in some of the cases (mostly
Zn7.5, Figure 2), the porosity of all the prepared materials decreased with increasing compacting
pressure and even decreased with the application of temperature during the material processing (CP-S
and HP). This fact can be attributed to a wide distribution of powder particle size, when the space
between large particles was filled up with the smaller ones. However, the porosity of materials from
Zn7.5 powder was substantially higher than in the case of material prepared from Zn150 powder,
which can be also observed from the fractographic evaluation of broken samples (Figures 5 and 6).
This can be explained by the larger range of the powder particle size measured for Zn150 powder (up
to 640 μm) compared to Zn7.5 powder (up to 20 μm) and the adequate arrangement of the particles
with different size in the compacted sample volume.
CP-S processing of Zn7.5 powder materials did not influence the material microstructure with
as much intensity as the increasing compression pressure did. Only the minor effect of the following
sintering of CP materials was observed by metallographic analysis (Figure 2). In the case of Zn150
powder materials, the following sintering of CP samples led to the increase of the porosity of materials
prepared at low pressures (Figure 3c). Due to the high sintering temperature of 400 ◦ C and high
material porosity (of CP samples), the powder particles reduced the surface energy and changed their
shape from irregular rod-shaped to spherical during sintering. At the same time, no dimensional
changes of prepared samples were observed, which would generally correspond if a solid-state
diffusion was the primary sintering mechanism [34]. The shape changes of particles and no shrinkage
could lead to pores growing between the particles.
The effect of HP processing is more evident on the Zn7.5 powder-based materials when
compared to the Zn150 powder materials. Elevated temperature during the material processing
resulted in improved plasticity and enhanced deformability of fine powder particles (Zn7.5 powder).
The particles’ deformation is observable on materials’ microstructures documented in Figure 2e,f.
Elevated temperature applied during material processing resulted in activation of more slip systems

182
Metals 2017, 7, 396

in the hexagonal close packed (HCP) structure (also characteristic for zinc) [34]. In the case of Zn150
powder materials, the elevated temperature of the processing had no evident influence on material
microstructure. In the case of smaller powder particles, the improved deformability of the material
due to the activation of more slip systems was much more significant compared to the larger powder
particles, which were deformable even at room temperature. Quite good deformability of large powder
particles was not significantly improved and remained the same as in the case of CP and CP-S materials
from the microstructural point of view, Figure 3.
There is only a small number of articles dealing with evaluation of the mechanical properties of
pure zinc [17,25–28]. From the mechanical properties point of view, materials in these studies were
characterized mostly by Vickers hardness and tensile tests. The Vickers hardness of experimental
materials prepared from Zn7.5 and Zn150 powders in the presented work was determined as
40–49 HV025 and 40–45 HV025, respectively. The values were higher in comparison to the cast pure zinc
of 30 HV given in [17]; however, the values were comparable with results of Vickers hardness testing
of hot extruded pure zinc of 44 HV5 given in [27]. Due to the obtained microstructure of materials
prepared by powder metallurgy, the characteristic mechanical properties should correlate more with the
properties of wrought materials than with cast materials. Considering the powder particles’ boundaries
as grain boundaries, the microstructure of the materials processed from Zn7.5 powder (knowing the
real particle size was up to 20 μm) is comparable with the fine recrystallized equi-axed grains with an
average grain size of 20 μm characteristic for pure extruded zinc analyzed in [27].
The resistance of prepared powder-based materials against the fracture during the three-point
bend test was (besides particles deformability) influenced mainly by the mechanical interlocking of
irregularities on the powder particle surfaces, which was promoted by plastic deformation during the
pressing. The bonding of particles was responsible for the mechanism of the bulk material fracture.
Because of the spherical shape of Zn7.5 powder particles, the flexural strength of compacted samples
showed lower values when compared to materials prepared from Zn150 powder. Larger surface area
and particles shape of Zn150 powder particles allowed easier contact of particles and interlocking of
the surface irregularities, which was even enhanced by the particles’ larger deformability compared to
the Zn7.5 powders, which was easily observed in the case of CP and CP-S materials, Figure 3.
The surface layer of corrosion products on particles’ surface had also a negative influence on
powder particles’ compaction during sintering and HP (processes performed at elevated temperature,
enhancing diffusion processes). Only the diffusion mechanisms contributed to sintering of particles
in the case of the absence of an external pressure [35]. The oxide layer on the powder particles could
affect the diffusion bonding of particles during sintering, and this can be seen in the case of CP-S
Zn150 powder material compacted under 100 MPa (Figure 3c). In this case, the observed porosity was
higher than as in the case of CP materials. Due to the surface oxide layer on the powder particles,
the sintering process could be affected by the diffusion delay. The role of the surface oxide layer in the
sintering of metal powder particles was defined as follows: Shifting the mechanisms of the sintering
process from a bulk-transport mechanism to that controlled by surface transport [34]. Even though
the content of oxygen was detected to be higher on the Zn7.5 powder particles, the powder particles’
low deformability seemed to have a larger influence on material compaction than the oxide layer on
the powder particles. Due to the low particle deformability and spherical shape, only limited areas of
individual particles were in direct contact, and the conditions for diffusion were worse compared to
the larger and more plane particles of Zn150 powder material. However, according to the experimental
observations, the influence of the oxide layer on the sinterability of zinc was minimal, and the presence
of an oxide layer on powder particles makes essentially no difference in the kinetics of sintering [34].
Even though the influence of processing parameters on Zn powder-based materials’ microstructure
is minimal, the influence of the processing on the samples’ bending properties was observed.
CP-S materials reached higher values of flexural strength for both powder particles sizes compared to
the CP materials. The influence was, however, only minor compared to the influence of the HP.

183
Metals 2017, 7, 396

In the case of Zn7.5 powder material, the CP samples prepared under 100, 200 and 300 MPa did
not reach the required handling strength necessary for three-point bend testing, and only samples
prepared under 400 and 500 MPa were tested. However, the values reached for the flexural strength
and displacement before fracture were very low (Figure 4a,c), which corresponds to low particle
deformation and the subsequent low bonding and high porosity of the material (Figure 2a,b).
The transgranular fracture mechanism and porosity present on the fracture surfaces support this
theory (Figure 5a,b). Sintering of the Zn7.5 powder material resulted in higher material handling
strength; however, the values of flexural strength reached were still low for materials prepared
under 100–300 MPa (Figure 4a). The fracture surface of the CP-S sample prepared under 200 MPa
corresponded to the microstructural observation and low bending characteristics of the material
reached (Figure 5c). Sintering materials prepared under higher pressures resulted in flexural strength
values higher than 40 MPa (41 and 76 MPa for 400 and 500 MPa pressure, respectively). However,
the displacement before the fracture was still very low and comparable to the one of CP samples
prepared under the same pressures (Figure 4c). HP of Zn7.5 powder materials resulted in a significant
increase of materials’ flexural strength, reaching maximal value of 322 MPa for material prepared under
400 MPa (Figure 4a). The positive influence of compacting pressure of HP-processed Zn7.5 powder
material was revealed by three-point bend test up to the compacting pressure of 400 MPa. Samples
prepared under 500 MPa reached lower values of flexural strength than the maximum, which can be
connected with the limited deformability of the material due to the HCP crystallographic structure
of zinc [36]. High compaction of the material during HP under 500 MPa resulted in strong particle
bonding, limiting the transgranular fracture mechanism. With increasing compacting pressure, a larger
amount of broken powder particles can be observed on the samples’ fracture surfaces (Figure 5e,f).
The cleavage mechanism playing a role in the material failure was responsible for the measured
decrease of the flexural strength (HP material processed under 500 MPa); however, displacement
before fracture still increased.
In the case of Zn150 powder materials, only the sample prepared by CP under 100 MPa did
not reach the adequate handling strength necessary for the three-point bend test. Due to the higher
deformability of larger powder particles compared to the Zn7.5 powder materials, CP and CP-S
samples reached higher values of flexural strength and displacement before fracture when compared
to the Zn7.5 samples. In the case of CP-S Zn150 powder materials, sintering had a significantly positive
influence on material bending properties (improvement of flexural strength approximately twice
compared to the CP materials). HP Zn150 powder materials reached even larger values of bending
properties than the CP-S materials; however, the improvement was not as significant as in the case of
Zn7.5 powder materials (Figure 4). The differences between the CP-S and HP Zn150 powder materials
are significant for materials prepared under 100–300 MPa; however, only a minor influence was
observed in the case of materials prepared under 400 and 500 MPa (Figure 4). This corresponds to the
microstructural observations, where only a minor influence on the materials microstructure applying
higher pressure (400–500 MPa) and temperature during HP was observed (Figure 3). The bonding
of particles of materials prepared under 300–500 MPa was comparable for CP-S and HP materials.
HP Zn150 powder materials’ fracture surfaces were similar for all of the applied pressures during
the preparation of materials, while differences in the CP-S materials’ fracture surfaces were observed
(Small pressures during materials processing resulted in intergranular failure, while the combination
of inter- and trans-granular fracture correlated with higher compacting pressures (Figure 6)).
Detailed fractographic analysis of the fracture surfaces of HP processed materials revealed the
mechanism of crack propagation (Figures 5f and 6f). Due to the tensile loading, microcracks through
individual powder particles were observed, characteristic of the cleavage facets in Figures 5f and 6f.
The final crack responsible for the sample’s failure followed the powder particles’ boundaries
connecting cleavage facets (cracked particles).

184
Metals 2017, 7, 396

5. Conclusions
Pure zinc materials prepared by methods of powder metallurgy were examined from the
microstructural and mechanical properties point of view. Bulk materials from zinc powder of a
mean particles size of 7.5 and 150 μm were prepared with three preparation methods: Cold pressing
(CP), cold pressing followed by sintering (CP-S) and hot pressing (HP). The obtained results show the
following:

(1) The deformability of zinc powder substantially depends on the size and shape of the powder
particles. Smaller spherical particles of Zn7.5 were observed to be only slightly deformable at
room temperature (CP), whereas large irregularly-shaped particles of Zn150 powder showed good
deformability resulting in low porosity of samples and good material compaction. The addition of
sintering (CP-S) and HP improved the deformability of the Zn7.5 powder; however, no significant
influence was observed in the case of Zn150 powder.
(2) The increase of compacting pressure was connected with the increase of the bending
characteristics of prepared Zn powder materials.
(3) The flexural strength of CP Zn150 powder materials was substantially higher when compared to
materials prepared from Zn7.5 powder. Low deformability of small-sized powder particles led
to the poor mechanical interlocking and low flexural strength of materials prepared from Zn7.5
powder compacted at room temperature.
(4) CP-S of Zn powder-based materials led to the significant improvement of the bending
characteristics of prepared materials. The maximal values of the flexural strength were measured
for CP-S materials prepared under 500 MPa; 76 MPa for Zn7.5 powder and 148 MPa for Zn150
powder materials.
(5) HP resulted in even more pronounced improvement of the bending characteristics of Zn
powder-based materials. The highest value reached of the flexural strength were measured
for Zn7.5 material prepared under 400 MPa (322 MPa) and for Zn150 material prepared under
100 and 200 MPa (203 MPa).
(6) The intercrystalline fracture mechanism was characteristic for CP materials prepared under
all of the applied pressures and CP-S materials prepared under pressures from 100–400 MPa.
The combination of inter- and trans-crystalline fracture mechanisms was characteristic for CP-S
materials prepared under 500 MPa and for HP materials.

Acknowledgments: This work was supported by Project No. LO1211, Materials Research Centre and by Project
No. LO1202, Center of New Technologies for Mechanical Engineering (NETME), plus the projects of the Ministry
of Education, Youth and Sports of the Czech Republic under the “National sustainability program”. We are also
grateful to Pavla Puškášová and Jakub Puškáš for the professional English editing service.
Author Contributions: Pavel Doležal conceived of and designed the experiments. Michaela Krystýnová
and Matěj Březina performed the experiments. Pavel Doležal, Michaela Krystýnová, Stanislava Fintová
and Matěj Březina analyzed the data. Pavel Doležal, Josef Zapletal and Jaromír Wasserbauer contributed
reagents/materials/analysis tools. Michaela Krystýnová and Stanislava Fintová wrote the paper.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Witte, F.; Hort, N.; Vogt, C.; Cohen, S.; Kainer, K.U.; Willumeit, R.; Feyerabend, F. Degradable biomaterials
based on magnesium corrosion. Curr. Opin. Solid State Mater. Sci. 2008, 12, 63–72. [CrossRef]
2. Tan, L.; Yu, X.; Wan, P.; Yang, K. Biodegradable Materials for Bone Repairs: A Review. J. Mater. Sci. Technol.
2013, 29, 503–513. [CrossRef]
3. Staiger, M.P.; Pietak, A.M.; Huadmai, J.; Dias, G. Magnesium and its alloys as orthopedic biomaterials:
A review. Biomaterials 2006, 27, 1728–1734. [CrossRef] [PubMed]
4. Heublein, B.; Rohde, R.; Kaese, V.; Niemeyer, M.; Hartung, W.; Haverich, A. Biocorrosion of magnesium
alloys: A new principle in cardiovascular implant technology? Heart 2003, 89, 651–656. [CrossRef] [PubMed]

185
Metals 2017, 7, 396

5. Waksman, R.; Pakala, R.; Kuchulakanti, P.K.; Baffour, R.; Hellinga, D.; Seabron, R.; Tio, F.O.; Wittchow, E.;
Hartwig, S.; Harder, C.; et al. Safety and efficacy of bioabsorbable magnesium alloy stents in porcine coronary
arteries. Catheter. Cardiovasc. Interv. 2006, 68, 607–617. [CrossRef] [PubMed]
6. Shuai, C.; Zhou, Y.; Yang, Y.; Feng, P.; Liu, L.; He, C.; Zhao, M.; Yang, S.; Gao, C.; Wu, P. Biodegradation
Resistance and Bioactivity of Hydroxyapatite Enhanced Mg-Zn Composites via Selective Laser Melting.
Materials 2017, 10, 307. [CrossRef] [PubMed]
7. Chen, J.; Wu, P.; Wang, Q.; Yang, Y.; Peng, S.; Zhou, Y.; Shuai, C.; Deng, Y. Influence of Alloying Treatment
and Rapid Solidification on the Degradation Behavior and Mechanical Properties of Mg. Metals 2016, 6, 259.
[CrossRef]
8. He, C.; Bin, S.; Wu, P.; Gao, C.; Feng, P.; Yang, Y.; Liu, L.; Zhou, Y.; Zhao, M.; Yang, S.; et al. Microstructure
Evolution and Biodegradation Behavior of Laser Rapid Solidified Mg–Al–Zn Alloy. Metals 2017, 7, 105. [CrossRef]
9. Gu, X.; Zheng, Y.; Cheng, Y.; Zhong, S.; Xi, T. In vitro corrosion and biocompatibility of binary magnesium
alloys. Biomaterials 2009, 30, 484–498. [CrossRef] [PubMed]
10. Ilich, J.Z.; Kerstetter, J.E. Nutrition in Bone Health Revisited: A Story beyond Calcium. J. Am. Coll. Nutr.
2000, 19, 715–737. [CrossRef] [PubMed]
11. Xu, L.; Yu, G.; Zhang, E.; Pan, F.; Yang, K. In vivo corrosion behavior of Mg-Mn-Zn alloy for bone implant
application. J. Biomed. Mater. Res. Part A 2007, 83, 703–711. [CrossRef] [PubMed]
12. Hänzi, A.C.; Gerber, I.; Schinhammer, M.; Löffler, J.F.; Uggowitzer, P.J. On the in vitro and in vivo degradation
performance and biological response of new biodegradable Mg–Y–Zn alloys. Acta Biomater. 2010, 6,
1824–1833. [CrossRef] [PubMed]
13. Zhang, E.; Yin, D.; Xu, L.; Yang, L.; Yang, K. Microstructure, mechanical and corrosion properties and
biocompatibility of Mg–Zn–Mn alloys for biomedical application. Mater. Sci. Eng. C 2009, 29, 987–993. [CrossRef]
14. Li, Z.; Gu, X.; Lou, S.; Zheng, Y. The development of binary Mg–Ca alloys for use as biodegradable materials
within bone. Biomaterials 2008, 29, 1329–1344. [CrossRef] [PubMed]
15. Gu, X.; Zheng, Y.; Zhong, S.; Xi, T.; Wang, J.; Wang, W. Corrosion of, and cellular responses to Mg–Zn–Ca
bulk metallic glasses. Biomaterials 2010, 31, 1093–1103. [CrossRef] [PubMed]
16. Zhang, S.; Zhang, X.; Zhao, C.; Li, J.; Song, Y.; Xie, C.; Tao, H.; Zhang, Y.; He, Y.; Jiang, Y. Research on an
Mg–Zn alloy as a degradable biomaterial. Acta Biomater. 2010, 6, 626–640. [CrossRef] [PubMed]
17. Vojtěch, D.; Kubásek, J.; Šerák, J.; Novák, P. Mechanical and corrosion properties of newly developed
biodegradable Zn-based alloys for bone fixation. Acta Biomater. 2011, 7, 3515–3522. [CrossRef] [PubMed]
18. Tapiero, H.; Tew, K.D. Trace elements in human physiology and pathology: Zinc and metallothioneins.
Biomed. Pharmacother. 2003, 57, 399–411. [CrossRef]
19. Aggett, P.J.; Harries, J.T. Current status of zinc in health and disease states. Arch. Dis. Child. 1979, 54, 909–917.
[CrossRef] [PubMed]
20. Fosmire, G.J. Zinc toxicity. Am. J. Clin. Nutr. 1990, 51, 225–227. [PubMed]
21. Bowen, P.K.; Drelich, J.; Goldman, J. Zinc Exhibits Ideal Physiological Corrosion Behavior for Bioabsorbable
Stents. Adv. Mater. 2013, 25, 2577–2582. [CrossRef] [PubMed]
22. Wang, Y.; Wei, M.; Gao, J.; Hu, J.; Zhang, Y. Corrosion process of pure magnesium in simulated body fluid.
Mater. Lett. 2008, 62, 2181–2184. [CrossRef]
23. Ollig, J.; Kloubert, V.; Weßels, I.; Haase, H.; Rink, L. Parameters Influencing Zinc in Experimental Systems in
Vivo and in Vitro. Metals 2016, 6, 71. [CrossRef]
24. Kirkland, N.T.; Lespagnol, J.; Birbilis, N.; Staiger, M.P. A survey of bio-corrosion rates of magnesium alloys.
Corros. Sci. 2010, 52, 287–291. [CrossRef]
25. ASM International. ASM Handbook, 10th ed.; ASM International: Materials Park, OH, USA, 2004.
26. Brandes, E.A.; Brook, G.B.; Smithells, C.J. Smithells Metals Reference Book, 7th ed.; Brandes, E.A., Brook, G.B.,
Eds.; Butterworth-Heinemann: Boston, MA, USA, 1998; Volume 22, p. 94.
27. Kubásek, J.; Vojtěch, D.; Jablonská, E.; Pospíšilová, I.; Lipov, J.; Ruml, T. Structure, mechanical characteristics
and in vitro degradation, cytotoxicity, genotoxicity and mutagenicity of novel biodegradable Zn–Mg alloys.
Mater. Sci. Eng. C 2016, 58, 24–35. [CrossRef] [PubMed]
28. Sikora-Jasinska, M.; Mostaed, E.; Mostaed, A.; Beanland, R.; Mantovani, D.; Vedani, M. Fabrication,
mechanical properties and in vitro degradation behavior of newly developed Zn Ag alloys for degradable
implant applications. Mater. Sci. Eng. C 2017, 77, 1170–1181. [CrossRef] [PubMed]

186
Metals 2017, 7, 396

29. Wang, X.; Lu, H.M.; Li, X.L.; Li, L.; Zheng, Y.F. Effect of cooling rate and composition on microstructures and
properties of Zn-Mg alloys. Trans. Nonferrous Met. Soc. China 2007, 17, 122–125.
30. Bowen, P.K.; Seitz, J.M.; Guillory, R.J.; Braykovich, J.P.; Zhao, S.; Goldman, J.; Drelich, J.W. Evaluation of
wrought Zn-Al alloys (1, 3, and 5 wt % Al) through mechanical and in vivo testing for stent applications.
J. Biomed. Mater. Res. Part B Appl. Biomater. 2017. [CrossRef] [PubMed]
31. Demirtas, M.; Purcek, G.; Yanar, H.; Zhang, Z.J.; Zhang, Z.F. Effect of chemical composition and grain size on
RT superplasticity of Zn-Al alloys processed by ECAP. Lett. Mater. 2015, 5, 328–334. [CrossRef]
32. Shaw, C.A.; Tomljenovic, L. Aluminum in the central nervous system (CNS): Toxicity in humans and animals,
vaccine adjuvants, and autoimmunity. Immunol. Res. 2013, 56, 304–316. [CrossRef] [PubMed]
33. Baba, K.; Hatada, R.; Flege, S.; Ensinger, W.; Shibata, Y.; Nakashima, J.; Sawase, T.; Morimura, T. Preparation
and antibacterial properties of Ag-containing diamond-like carbon films prepared by a combination of
magnetron sputtering and plasma source ion implantation. Vacuum 2013, 89, 179–184. [CrossRef]
34. Munir, Z.A. Analytical treatment of the role of surface oxide layers in the sintering of metals. J. Mater. Sci.
1979, 14, 2733–2740. [CrossRef]
35. Nassef, A.; El-Garaihy, W.; El-Hadek, M. Mechanical and Corrosion Behavior of Al-Zn-Cr Family Alloys.
Metals 2017, 7, 171. [CrossRef]
36. Prasad, Y.V.R.K.; Sasidhara, S. Hot Working Guide: A Compendium of Processing Maps; ASM International:
Materials Park, OH, USA, 1997; p. 496.

© 2017 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

187
MDPI
St. Alban-Anlage 66
4052 Basel
Switzerland
Tel. +41 61 683 77 34
Fax +41 61 302 89 18
www.mdpi.com

Metals Editorial Office


E-mail: metals@mdpi.com
www.mdpi.com/journal/metals
MDPI
St. Alban-Anlage 66
4052 Basel
Switzerland
Tel: +41 61 683 77 34
Fax: +41 61 302 89 18
www.mdpi.com ISBN 978-3-03897-387-4

Das könnte Ihnen auch gefallen